This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Non-unitary Wightman CFTs and non-unitary vertex algebras

Sebastiano Carpi [email protected] Dipartimento di Matematica, Università di Roma Tor Vergata,
Via della Ricerca Scientifica 1, I-00133 Roma, Italy
Christopher Raymond [email protected] Department of Mathematics and Statistics, Université Laval, Quebec, Canada, G1V 0A6 Yoh Tanimoto [email protected] Dipartimento di Matematica, Università di Roma Tor Vergata,
Via della Ricerca Scientifica 1, I-00133 Roma, Italy
James E.​ Tener [email protected] Mathematical Sciences Institute, Australian National University,
Canberra, ACT 2600, Australia
Abstract

We give an equivalence of categories between: (i) Möbius vertex algebras which are equipped with a choice of generating family of quasiprimary vectors, and (ii) (not-necessarily-unitary) Möbius-covariant Wightman conformal field theories on the unit circle. We do not impose any technical restrictions on the theories considered (such as finite-dimensional conformal weight spaces or simplicity), yielding the most general equivalence between these two axiomatizations of two-dimensional chiral conformal field theory. This provides new opportunities to study non-unitary vertex algebras using the lens of algebraic conformal field theory and operator algebras, which we demonstrate by establishing a non-unitary version of the Reeh-Schlieder theorem.

1 Introduction

It is a fundamental mathematical challenge to establish a rigorous axiomatization of quantum field theory (QFT), and in general this problem remains wide open except in very specialized contexts. In recent years, axiomatic QFT has received particular attention in the context of two-dimensional chiral conformal field theories (CFTs), as these theories are sufficiently structured to enable a rigorous mathematical treatment while at the same time exhibiting a wide variety of mathematical connections (to operator algebras and subfactors, to representation theory and modular tensor categories, to vector-valued modular forms, and to many other areas). There are many proposed axiomatizations of two-dimensional chiral CFTs, each of which captures different aspects of the physical theory, and none of which have been rigorously demonstrated to provide a complete description of CFT. It is conjectured that these different axiomatizations are essentially equivalent, and there have been recent breakthroughs in comparing different axiomatizations under certain technical hypotheses [CKLW18, Ten19a].

In this article we demonstrate the equivalence of two well-known axiomatizations of two-dimensional chiral CFTs. We establish this equivalence in the most general way possible, without any reliance on auxiliary technical hypotheses or restrictions on the models under consideration, such as the existence of an invariant inner product (“unitarity”). Proving equivalences at this level of generality has largely been viewed as an aspirational (but not necessarily feasible) goal of axiomatic QFT, which we achieve here through a detailed analysis of the mathematical structures in question.

The first axiomatization that we consider is the non-unitary version of the (bosonic) Wightman axioms on the unit circle S1S^{1}\subset{\mathbb{C}}, with Möbius symmetry (i.e.​ symmetry group Mo¨b:=PSU(1,1)\operatorname{M\ddot{o}b}:=\mathrm{PSU}(1,1), the holomorphic automorphisms of the unit disk). The key data of such a theory is a collection {\mathcal{F}} of operator-valued distributions (or Wightman fields) acting on a common invariant vector space of states 𝒟{\mathcal{D}}, along with a compatible positive-energy representation of Mo¨b\operatorname{M\ddot{o}b}.

The second axiomatization that we consider is ({\mathbb{N}}-graded, bosonic) Möbius vertex algebras. These are vertex algebras graded by non-negative integer conformal dimensions, with symmetry given by the complexified Lie algebra 𝔰𝔲(1,1)𝔰𝔩2()\mathfrak{su}(1,1)_{\mathbb{C}}\cong\mathfrak{sl}_{2}({\mathbb{C}}).

We prove the following main result.

Main Result.

There is a natural equivalence of categories between non-unitary Möbius-covariant Wightman conformal field theories on S1S^{1} and Möbius vertex algebras equipped with a family of quasiprimary generators.

Our result does not require unitarity or the existence of an invariant bilinear form, and we do not require that the homogeneous subspaces for the grading by conformal dimensions be finite-dimensional. There are many important examples of CFTs arising in mathematical and theoretical physics which require this level of generality, and in particular non-unitarity arises from the CFT-approach to classical critical phenomena, and from string theory. Specific examples include the non-unitary Virasoro minimal models, affine vertex algebras at non-critical level (both universal and simple quotient), bosonic {\mathbb{N}}-graded affine W-algebras (again universal and simple quotient), and the βγ\beta\gamma-ghost vertex algebra with central charge c=2c=2 (along with other “AA-graded” vertex algebras which arise in logarithmic conformal field theory). As a result of our theorem there are canonical Wightman CFTs associated to these models, which demonstrates significant functional analytic regularity that is not otherwise apparent.

The constructions going from Wightman CFTs to vertex algebras and back are given in Sections 3.1 and 3.2, respectively. These are shown to give an equivalence of categories in Section 3.3. The vertex algebra 𝒱{\mathcal{V}} associated to a Wightman CFT with domain 𝒟{\mathcal{D}} is constructed as a certain subspace 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}}. Conversely, the Wightman CFT associated to 𝒱{\mathcal{V}} is constructed as an extension 𝒱𝒟𝒱(n){\mathcal{V}}\subset{\mathcal{D}}\subset\prod{\mathcal{V}}(n). We note that at this level of generality (allowing each weight space 𝒱(n){\mathcal{V}}(n) to be infinite-dimensional) it is not a priori clear that there is a single Wightman CFT for each vertex algebra, and it seems plausible that there could be families of ‘Wightman completions’ of a single vertex algebra. However, as a consequence of our result, there is indeed a unique Wightman CFT for each Möbius vertex algebra.

A very useful and inspiring heuristic discussion on the connection between Wightman CFTs and vertex algebras can be found in [Kac98, §1.2]. However, the arguments given there do not appear to be aimed to give precise mathematical details on this connection. More recently, three of the present authors gave a rigorous proof that unitary Möbius vertex algebras were equivalent to unitary Wightman CFTs possessing an additional analytic property called uniformly bounded order, provided that the homogeneous subspaces for the grading by conformal dimensions were finite-dimensional [RTT22]. The present article generalizes the previous result to possibly non-unitary theories, also dropping the requirements of uniformly bounded order and finite-dimensional weight spaces. As the techniques historically used to study Wightman theories involve careful analysis of the norm topology on the space of states, there is significant new work required to generalize our previous results to the non-unitary setting.

We also demonstrate in Section 4 that the correspondence constructed in this article is compatible with invariant bilinear forms, invariant sesquilinear forms, and invariant inner products.

Theorem.

Let 𝒟{\mathcal{D}} be a Möbius-covariant Wightman CFT and let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} be the associated Möbius vertex algebra. Then every invariant inner product (unitary structure) on 𝒟{\mathcal{D}} restricts to an invariant inner product on 𝒱{\mathcal{V}}, and conversely every unitary structure on 𝒱{\mathcal{V}} uniquely extends to one on 𝒟{\mathcal{D}}. The same holds for invariant sesquilinear forms (involutive structures) and invariant bilinear forms.

We are left with a striking and clear correspondence between two well-known axiomatizations of two-dimensional chiral conformal field theory, without any reliance on additional technical hypotheses. We are motivated in part by the possibility to provide such an equivalence, which is not often possible in the wild landscape of axiomatic quantum field theory. We are also motivated by intriguing links between non-unitary conformal field theories and the unitary world of algebraic conformal field theory. Given a Wightman CFT on S1S^{1} and an interval IS1I\subset S^{1}, consider the algebra 𝒫(I){\mathcal{P}}(I) generated by Wightman fields φ(f)\varphi(f) smeared by test functions ff supported in the interval II. Such Wightman nets of algebras have been studied in the context of unitary quantum field theories [SW64], and there is a substantial effort underway to understand the relationship between unitary vertex algebras, unitary Wightman nets, and the usual nets of algebras of bounded observables (i.e.​ conformal nets) studied in algebraic conformal field theory [CKLW18, Ten19a]. On the other hand, as a result of our present work, there exists a Wightman net for every Möbius vertex algebra, including non-unitary ones. Such nets could give an avenue to apply methods generally used in the unitary framework of algebraic quantum field theory in the more general setting of non-unitary models. Previously such links have been probed only at the level of categories of representations [EG17]. As a first demonstration of the potential of this approach we prove a version of the Reeh-Schlieder Theorem (regarding the cyclic and separating property of the vacuum vector) for non-unitary theories in Appendix A.

Finally, there is strong motivation to understand functional analytic aspects of non-unitary vertex algebras as a part of studying links between algebraic and geometric aspects of the theory, as in [Hua99, Hua03]. More recently, analytic considerations of non-unitary vertex algebras have played a key role in the study of conformal blocks [Gui24a, Gui24b, GZ23], and such considerations also feature in the construction of functorial CFTs in the sense of Segal [Seg04].

In future work, it would be interesting to relate modules for vertex algebras to representations of the corresponding Wightman nets, which would fit into the broad program underway in the unitary setting to relate vertex algebra modules to representations in algebraic conformal field theory [Ten19b, Ten24, Gui21, Gui20, CWX]. Such relations should enable further correspondences between full two-dimensional conformal field theories in various approaches, cf.​ [Mor23, AGT23, AMT24].

Acknowledgements

S.C. and Y.T acknowledge support from the GNAMPA-INDAM project Operator algebras and infinite quantum systems, CUP E53C23001670001 and from the MIUR Excellence Department Project MatMod@TOV awarded to the Department of Mathematics, University of Rome “Tor Vergata”, CUP E83C23000330006. C.R. and J.T. were supported by ARC Discovery Project DP200100067, “Physical realisation of enriched quantum symmetries”.

2 Preliminaries on Wightman CFTs and Möbius vertex algebras

An operator-valued distribution on the unit circle S1S^{1} with domain a vector space 𝒟\mathcal{D} is a linear map

φ:C(S1)(𝒟),\varphi:C^{\infty}(S^{1})\to{\mathcal{L}}({\mathcal{D}}),

where (𝒟){\mathcal{L}}({\mathcal{D}}) is the space of linear operators on 𝒟{\mathcal{D}}. In this article, we will typically study operator-valued distributions whose domain 𝒟{\mathcal{D}} is infinite-dimensional, and we will require some topological considerations with respect to the action of sets of such distributions on 𝒟{\mathcal{D}}.

If {\mathcal{F}} is a set of operator-valued distributions on S1S^{1} with a common domain 𝒟{\mathcal{D}}, then a linear functional λ:𝒟\lambda:{\mathcal{D}}\to{\mathbb{C}} is called compatible with {\mathcal{F}} if the multilinear maps C(S1)kC^{\infty}(S^{1})^{k}\to{\mathbb{C}} given by

(f1,,fk)λ(φ1(f1)φk(fk)Φ)(f_{1},\ldots,f_{k})\mapsto\lambda\big{(}\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi\big{)} (2.1)

are continuous in the fjf_{j} for all φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}}. Note that multilinear forms C(S1)kC^{\infty}(S^{1})^{k}\to{\mathbb{C}} are separately continuous if and only if they are jointly continuous since C(S1)C^{\infty}(S^{1}) is a Fréchet space [Trè67, Cor. §34.2]. We write 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} for the space of all linear functionals compatible with {\mathcal{F}}. Recall that a set 𝒳{\mathcal{X}} of linear functionals on 𝒟{\mathcal{D}} is said to separate points if for every non-zero Φ𝒟\Phi\in{\mathcal{D}} there is a λ𝒳\lambda\in{\mathcal{X}} such that λ(Φ)0\lambda(\Phi)\neq 0.

Definition 2.1.

A set {\mathcal{F}} of operator-valued distributions with domain 𝒟{\mathcal{D}} acts regularly if 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} separates points.

If we imagine that {\mathcal{F}} consists of a family of Wightman fields (i.e.​ the operators φ(f)\varphi(f) are smeared quantum fields), then it is natural that expectations λ:𝒟\lambda:{\mathcal{D}}\to{\mathbb{C}} should have the property that expressions (2.1) are continuous in the smearing functions fjf_{j}. Thus, the condition of regularity serves to exclude certain nonphysical actions that have the property that expectations cannot distinguish states. The following example illustrates the pathological behavior of nonregular actions.

Example 2.2.

Let 𝒟=T(C(S1))=k=0C(S1)k{\mathcal{D}}=T(C^{\infty}(S^{1}))=\bigoplus_{k=0}^{\infty}C^{\infty}(S^{1})^{\otimes k} be the tensor algebra, and for fC(S1)f\in C^{\infty}(S^{1}) let φ(f)(𝒟)\varphi(f)\in{\mathcal{L}}({\mathcal{D}}) be the operation of left-multiplication by ff in 𝒟{\mathcal{D}}. The space 𝒟{\mathcal{D}} carries a regular action of ={φ}{\mathcal{F}}=\{\varphi\}. Let 𝒟{\mathcal{I}}\subsetneq{\mathcal{D}} be the left ideal generated by trigonometric polynomials [z±1]C(S1){\mathbb{C}}[z^{\pm 1}]\subset C^{\infty}(S^{1}). Let 𝒟~=𝒟/\tilde{\mathcal{D}}={\mathcal{D}}/{\mathcal{I}}, and observe that for each fC(S1)f\in C^{\infty}(S^{1}) the action of φ(f)\varphi(f) descends to an operator φ~(f)(𝒟~)\tilde{\varphi}(f)\in{\mathcal{L}}(\tilde{\mathcal{D}}). The action of ~={φ~}\tilde{\mathcal{F}}=\{\tilde{\varphi}\} on 𝒟~\tilde{\mathcal{D}} is not regular. Let Ω𝒟\Omega\in{\mathcal{D}} be the unit of the tensor algebra, and let Ω~𝒟~\tilde{\Omega}\in\tilde{\mathcal{D}} be its image under the canonical projection. For any f[z±1]f\in{\mathbb{C}}[z^{\pm 1}] we have φ~(f)Ω~=0\tilde{\varphi}(f)\tilde{\Omega}=0, and thus any λ𝒟~~\lambda\in\tilde{\mathcal{D}}_{\tilde{\mathcal{F}}}^{*} vanishes on φ~(f)Ω~\tilde{\varphi}(f)\tilde{\Omega} for any fC(S1)f\in C^{\infty}(S^{1}). In particular, if fC(S1)[z±1]f\in C^{\infty}(S^{1})\setminus{\mathbb{C}}[z^{\pm 1}], then φ~(f)Ω~\tilde{\varphi}(f)\tilde{\Omega} is non-zero but lies in the kernel of all λ𝒟~~\lambda\in\tilde{\mathcal{D}}_{\tilde{\mathcal{F}}}^{*}.

Remark 2.3.

A non-regular action of {\mathcal{F}} on 𝒟{\mathcal{D}} descends to a regular action on the quotient 𝒟/𝒟0{\mathcal{D}}/{\mathcal{D}}_{0}, where 𝒟0=λ𝒟kerλ{\mathcal{D}}_{0}=\bigcap_{\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}}\ker\lambda.

Let enC(S1)e_{n}\in C^{\infty}(S^{1}) be the function en(z)=zne_{n}(z)=z^{n}. The condition that {\mathcal{F}} acts regularly on 𝒟{\mathcal{D}} ensures that the operators φ(f)\varphi(f) are determined by the modes φ(en)\varphi(e_{n}) in a certain sense that we will make precise below. This is in contrast with Example 2.2, in which φ~(en)Ω~=0\tilde{\varphi}(e_{n})\tilde{\Omega}=0 for all nn but φ~(f)Ω~0\tilde{\varphi}(f)\tilde{\Omega}\neq 0 for some fC(S1)f\in C^{\infty}(S^{1}).

We now introduce certain topologies on 𝒟{\mathcal{D}} associated with the action of {\mathcal{F}}. We assume here that the reader is familiar with (or indifferent to) the fundamentals of topological and locally convex vector spaces, and defer the relevant background and additional details to Appendix B.

Definition 2.4.

Given a family {\mathcal{F}} of operator-valued distributions on S1S^{1} with domain 𝒟{\mathcal{D}}, the {\mathcal{F}}-weak topology on 𝒟{\mathcal{D}} is the weak topology induced by the linear functionals 𝒟{\mathcal{D}}_{\mathcal{F}}^{*}. That is, the {\mathcal{F}}-weak topology is the coarsest topology such that every λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*} is continuous.

For a topological vector space XX, a map T:X𝒟T:X\to{\mathcal{D}} is {\mathcal{F}}-weakly continuous precisely when λT\lambda\circ T is continuous for all λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}. The family {\mathcal{F}} acts regularly precisely when the {\mathcal{F}}-weak topology is Hausdorff. We will see in Lemma 2.8 below that for φ\varphi\in{\mathcal{F}} the expressions φ(f)Φ\varphi(f)\Phi are continuous in fC(S1)f\in C^{\infty}(S^{1}) when 𝒟{\mathcal{D}} is given the {\mathcal{F}}-weak topology, so indeed φ(f)\varphi(f) is determined by the modes φ(en)\varphi(e_{n}) when {\mathcal{F}} acts regularly.

There is a second natural topology on 𝒟{\mathcal{D}} associated with the action of {\mathcal{F}}. For φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}}, let

Sφ1,,φk,Φ:C(S1)k𝒟S_{\varphi_{1},\ldots,\varphi_{k},\Phi}:C^{\infty}(S^{1})^{\otimes k}\to{\mathcal{D}}

be the linear map

Sφ1,,φk,Φ(f1fk)=φ1(f1)φk(fk)Φ.S_{\varphi_{1},\ldots,\varphi_{k},\Phi}(f_{1}\otimes\cdots\otimes f_{k})=\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi.

We equip the algebraic tensor product C(S1)kC^{\infty}(S^{1})^{\otimes k} with the projective topology, for which continuous linear maps C(S1)kXC^{\infty}(S^{1})^{\otimes k}\to X correspond to continuous multilinear maps (see Appendix B).

Definition 2.5.

Given a family of {\mathcal{F}} of operator-valued distributions on S1S^{1} with domain 𝒟{\mathcal{D}}, the {\mathcal{F}}-strong topology on 𝒟{\mathcal{D}} is the colimit (or final) locally convex topology induced by the maps Sφ1,,φk,ΦS_{\varphi_{1},\ldots,\varphi_{k},\Phi} for φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}}. That is, the {\mathcal{F}}-strong topology is the finest locally convex topology such that the maps Sφ1,,φk,ΦS_{\varphi_{1},\ldots,\varphi_{k},\Phi} are continuous.

Equivalently, the {\mathcal{F}}-strong topology is the finest locally convex topology on 𝒟{\mathcal{D}} such that expressions φ1(f1)φk(fk)Φ\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi are continuous in the functions fjf_{j} (jointly, or equivalently separately by [Trè67, Cor. §34.2]).

Remark 2.6.

For a locally convex space XX, a linear map T:𝒟XT:{\mathcal{D}}\to X is continuous precisely when TSφ1,,φk,ΦT\circ S_{\varphi_{1},\ldots,\varphi_{k},\Phi} is continuous for all φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}} [NB11, Thm. 12.2.2]. In particular, a linear functional λ:𝒟\lambda:{\mathcal{D}}\to{\mathbb{C}} is {\mathcal{F}}-strongly continuous if and only if λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}, and so the weak topology on 𝒟{\mathcal{D}} induced by the space of {\mathcal{F}}-strong continuous linear functionals is precisely the {\mathcal{F}}-weak topology.

We now have the following alternate characterizations of the regularity of an action of {\mathcal{F}} on 𝒟{\mathcal{D}}.

Lemma 2.7.

Let {\mathcal{F}} be a set of operator-valued distributions on S1S^{1} with domain a vector space 𝒟{\mathcal{D}}. Then the following are equivalent.

  1. i)

    {\mathcal{F}} acts regularly on 𝒟{\mathcal{D}}, i.e.​ 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} separates points.

  2. ii)

    The {\mathcal{F}}-weak topology on 𝒟{\mathcal{D}} is Hausdorff.

  3. iii)

    The {\mathcal{F}}-strong topology on 𝒟{\mathcal{D}} is Hausdorff.

  4. iv)

    There exists a locally convex Hausdorff topology on 𝒟{\mathcal{D}} such that the maps

    (f1,,fk)φ1(f1)φk(fk)Φ(f_{1},\ldots,f_{k})\mapsto\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi

    are continuous C(S1)k𝒟C^{\infty}(S^{1})^{k}\to{\mathcal{D}} for all φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}}.

Proof.

As noted above, the implication (i) \implies (ii) follows immediately from the definitions of regularity and the {\mathcal{F}}-weak topology. The identity map 𝒟𝒟{\mathcal{D}}\to{\mathcal{D}} is continuous from the {\mathcal{F}}-weak topology to the {\mathcal{F}}-strong topology, and thus (ii) \implies (iii). The {\mathcal{F}}-strong topology is locally convex by definition, and thus (iii) \implies (iv) is tautological. Finally, if τ\tau is a locally convex Hausdorff topology on 𝒟{\mathcal{D}} as in (iv), then we have an inclusion of continuous duals (𝒟,τ)𝒟({\mathcal{D}},\tau)^{*}\subset{\mathcal{D}}_{\mathcal{F}}^{*}. By the Hahn-Banach theorem (𝒟,τ)({\mathcal{D}},\tau)^{*} separates points [NB11, Thm. 7.7.7]. Hence so does 𝒟{\mathcal{D}}_{\mathcal{F}}^{*}, and the action of {\mathcal{F}} is regular. ∎

Both the {\mathcal{F}}-strong and {\mathcal{F}}-weak topologies are quite natural, and so it is not surprising that the fields {\mathcal{F}} act continuously when 𝒟{\mathcal{D}} is given one of these topologies.

Lemma 2.8.

Let {\mathcal{F}} be a set of operator-valued distributions on S1S^{1} acting regularly with domain 𝒟{\mathcal{D}} equipped with the {\mathcal{F}}-strong topology. Then for φ\varphi\in{\mathcal{F}} the natural map φ:C(S1)×𝒟𝒟\varphi:C^{\infty}(S^{1})\times{\mathcal{D}}\to{\mathcal{D}} is separately continuous. The same holds if 𝒟{\mathcal{D}} is equipped with the {\mathcal{F}}-weak topology.

Proof.

First, fix fC(S1)f\in C^{\infty}(S^{1}). For any φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}}, the expression

φ(f)φ1(f1)φk(fk)Φ\varphi(f)\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi

is continuous in the fkf_{k} by the definition of the {\mathcal{F}}-strong topology, and thus φ(f):𝒟𝒟\varphi(f):{\mathcal{D}}\to{\mathcal{D}} is continuous by Remark 2.6. Similarly, for fixed Φ𝒟\Phi\in{\mathcal{D}} expressions φ(f)Φ\varphi(f)\Phi are continuous in ff, and we conclude that φ\varphi is separately continuous.

Now consider if 𝒟{\mathcal{D}} is given the {\mathcal{F}}-weak topology. For λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*} the expression λ(φ(f)Φ)\lambda(\varphi(f)\Phi) is evidently continuous in ff, and it just remains to show that φ(f)\varphi(f) acts continuously on (𝒟,-weak)({\mathcal{D}},{\mathcal{F}}\mbox{-weak}). Let 𝒟{\mathcal{D}}^{\sharp} be the algebraic dual of 𝒟{\mathcal{D}}, and let φ(f):𝒟𝒟\varphi(f)^{*}:{\mathcal{D}}^{\sharp}\to{\mathcal{D}}^{\sharp} be the adjoint action. If λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*} then

(φ(f)λ)(φ1(f1)φk(fk)Φ)=λ(φ(f)φ1(f1)φk(fk)Φ)(\varphi(f)^{*}\lambda)(\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi)=\lambda(\varphi(f)\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi) (2.2)

is continuous in the functions fjf_{j}, so φ(f)\varphi(f)^{*} leaves 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} invariant. Thus if Φn\Phi_{n} is a net in 𝒟{\mathcal{D}} converging {\mathcal{F}}-weakly to Φ\Phi then

λ(φ(f)Φn)=(φ(f)λ)Φn(φ(f)λ)Φ=λ(φ(f)Φ).\lambda(\varphi(f)\Phi_{n})=(\varphi(f)^{*}\lambda)\Phi_{n}\to(\varphi(f)^{*}\lambda)\Phi=\lambda(\varphi(f)\Phi).

Hence φ(f)Φn\varphi(f)\Phi_{n} converges {\mathcal{F}}-weakly to φ(f)Φ\varphi(f)\Phi and φ(f)\varphi(f) acts continuously on 𝒟{\mathcal{D}}. ∎

We now assume our vector space 𝒟{\mathcal{D}} is equipped with a family {\mathcal{F}} of operator-valued distributions on S1S^{1} as well as a representation U:Mo¨bEnd(𝒟)U:\operatorname{M\ddot{o}b}\to\operatorname{End}({\mathcal{D}}), where Mo¨b=PSU(1,1)\operatorname{M\ddot{o}b}=\mathrm{PSU}(1,1) is the group of holomorphic automorphisms of the closed unit disk. As in [CKLW18, §6], for γMo¨b\gamma\in\operatorname{M\ddot{o}b} we denote by XγC(S1)X_{\gamma}\in C^{\infty}(S^{1}) the function

Xγ(eiϑ)=iddϑlog(γ(eiϑ)),X_{\gamma}({\mathrm{e}}^{{\mathrm{i}}\vartheta})=-{\mathrm{i}}\frac{d}{d\vartheta}\log(\gamma({\mathrm{e}}^{{\mathrm{i}}\vartheta})),

which takes positive real values since γ\gamma is an orientation-preserving diffeomorphism of S1S^{1}. For fC(S1)f\in C^{\infty}(S^{1}) and d0d\in{\mathbb{Z}}_{\geq 0} we denote by βd(γ)fC(S1)\beta_{d}(\gamma)f\in C^{\infty}(S^{1}) the function

(βd(γ)f)(z)=(Xγ(γ1(z)))d1f(γ1(z)).(\beta_{d}(\gamma)f)(z)=(X_{\gamma}(\gamma^{-1}(z)))^{d-1}f(\gamma^{-1}(z)). (2.3)

An operator-valued distribution with domain 𝒟{\mathcal{D}} is called Möbius-covariant with conformal dimension dd under the representation UU if for every γMo¨b\gamma\in\operatorname{M\ddot{o}b} and every fC(S1)f\in C^{\infty}(S^{1}) we have

U(γ)φ(f)U(γ)1=φ(βd(γ)f)U(\gamma)\varphi(f)U(\gamma)^{-1}=\varphi(\beta_{d}(\gamma)f)

as endomorphisms of 𝒟{\mathcal{D}}. We say that a vector Φ𝒟\Phi\in{\mathcal{D}} has conformal dimension dd\in{\mathbb{Z}} if U(Rϑ)Φ=eidϑΦU(R_{\vartheta})\Phi={\mathrm{e}}^{{\mathrm{i}}d\vartheta}\Phi for all rotations RϑMo¨bR_{\vartheta}\in\operatorname{M\ddot{o}b}.

We now present the not-necessarily-unitary version of the Wightman axioms for two-dimensional chiral conformal field theories on the circle S1S^{1}. Historically, the Wightman axioms have been closely entwined with unitary theories, where the space of states 𝒟{\mathcal{D}} possess an appropriate inner product. Non-unitary versions of the Wightman axioms have also appeared in various contexts such as the mathematical description of gauge fields, see e.g.​ [Str93, §6.4]. In this article we will generally refer to the non-unitary theories in question simply as Wightman conformal field theories for the sake of brevity.

Definition 2.9.

Let 𝒟{\mathcal{D}} be a vector space equipped with a representation UU of Mo¨b\operatorname{M\ddot{o}b} and a choice of non-zero vector Ω𝒟\Omega\in{\mathcal{D}}. Let \mathcal{F} be a set of operator-valued distributions on S1S^{1} acting regularly on their common domain 𝒟\mathcal{D}. This data forms a (not-necessarily-unitary) Möbius-covariant Wightman CFT on S1S^{1} if they satisfy the following axioms:

  1. (W1)

    Möbius covariance: For each φ\varphi\in\mathcal{F} there is d0d\in\mathbb{Z}_{\geq 0} such that φ\varphi is Möbius-covariant with conformal dimension dd under the representation UU.

  2. (W2)

    Locality: If ff and gg have disjoint supports, then φ1(f)\varphi_{1}(f) and φ2(g)\varphi_{2}(g) commute for any pair φ1,φ2\varphi_{1},\varphi_{2}\in\mathcal{F}.

  3. (W3)

    Spectrum condition: If Φ𝒟\Phi\in{\mathcal{D}} has conformal dimension d<0d<0 then Φ=0\Phi=0.

  4. (W4)

    Vacuum: The vector Ω\Omega is invariant under UU, and 𝒟{\mathcal{D}} is spanned by vectors of the form φ1(f1)φk(fk)Ω\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega.

A Möbius-covariant Wightman CFT is a quadruple (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega), but we will frequently refer to the family {\mathcal{F}} of fields or the domain 𝒟{\mathcal{D}} as a (Möbius-covariant) Wightman CFT when the remaining data is clear from context.

Let ej(z)=zjC(S1)e_{j}(z)=z^{j}\in C^{\infty}(S^{1}), and let

𝒱(n)=span{φ1(ej1)φk(ejk)Ω|j1++jk=n}.{\mathcal{V}}(n)=\operatorname{span}\{\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega\,|\,j_{1}+\cdots+j_{k}=-n\}.

By Möbius covariance (W1) the vectors in 𝒱{\mathcal{V}} have conformal dimension nn, or in other words U(Rϑ)U(R_{\vartheta}) acts on 𝒱(n){\mathcal{V}}(n) as multiplication by einϑ{\mathrm{e}}^{{\mathrm{i}}n\vartheta}.

Lemma 2.10.

Let ({\mathcal{F}},𝒟{\mathcal{D}},UU,Ω\Omega) be a Möbius-covariant Wightman CFT, and suppose 𝒟{\mathcal{D}} is equipped with the {\mathcal{F}}-strong topology.

  1. i)

    The map U:Mo¨b×𝒟𝒟U:\operatorname{M\ddot{o}b}\times{\mathcal{D}}\to{\mathcal{D}} is separately continuous.

  2. ii)

    n0𝒱(n)\bigoplus_{n\geq 0}{\mathcal{V}}(n) is dense in 𝒟{\mathcal{D}}.

The same holds for the {\mathcal{F}}-weak topology.

Proof.

First consider the {\mathcal{F}}-strong topology. Fix Φ𝒟\Phi\in{\mathcal{D}}, and we will show that U(γ)ΦU(\gamma)\Phi is continuous in γ\gamma. By the vacuum axiom we may assume without loss of generality that Φ=φ1(f1)φk(fk)Ω\Phi=\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega. We have

U(γ)φ1(f1)φk(fk)Ω=φ1(βd1(γ)f1)φk(βdk(γ)fk)Ω.U(\gamma)\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega=\varphi_{1}(\beta_{d_{1}}(\gamma)f_{1})\cdots\varphi_{k}(\beta_{d_{k}}(\gamma)f_{k})\Omega.

The smooth function βd(γ)f\beta_{d}(\gamma)f depends continuously on γ\gamma, and thus U(γ)φ1(f1)φk(fk)ΩU(\gamma)\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega depends continuously on γ\gamma as well. Now consider a fixed γMo¨b\gamma\in\operatorname{M\ddot{o}b}, and by the same argument U(γ)φ1(f1)φk(fk)ΩU(\gamma)\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega depends continuously on the fjf_{j}. Hence by Remark 2.6 and the vacuum axiom U(γ)ΦU(\gamma)\Phi depends continuously on Φ\Phi, proving (i). The argument is similar for the {\mathcal{F}}-weak topology.

For (ii), note that n𝒱(n)\bigoplus_{n\in{\mathbb{Z}}}{\mathcal{V}}(n) is dense since Laurent polynomials are dense in C(S1)C^{\infty}(S^{1}), and 𝒱(n)=0{\mathcal{V}}(n)=0 for n<0n<0 by the spectrum condition. ∎

Let 𝒟𝒱(n){\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n)^{*} denote the space of linear functionals λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*} such that λ|𝒱(m)=0\lambda|_{{\mathcal{V}}(m)}=0 when mnm\neq n. The following technical observations will be essential in constructing vertex algebras from Wightman CFTs.

Lemma 2.11.

Suppose that ({\mathcal{F}},𝒟{\mathcal{D}},UU,Ω\Omega) satisfies all of the axioms of a Möbius-covariant Wightman CFT except perhaps for the spectrum condition.

  1. i)

    For nn\in{\mathbb{Z}}, 𝒟𝒱(n){\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n)^{*} separates points in 𝒱(n){\mathcal{V}}(n).

  2. ii)

    If n0𝒱(n)\bigoplus_{n\geq 0}{\mathcal{V}}(n) is {\mathcal{F}}-strongly (or {\mathcal{F}}-weakly) dense then the spectrum condition holds.

Proof.

First we prove (i). From the definition of a Wightman CFT, 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} separates points so given v𝒱(n)v\in{\mathcal{V}}(n) we may choose λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*} such that λ(v)0\lambda(v)\neq 0. For z=eiϑz={\mathrm{e}}^{{\mathrm{i}}\vartheta}, let rz=U(Rϑ)Mo¨br_{z}=U(R_{\vartheta})\in\operatorname{M\ddot{o}b} be rotation by zS1z\in S^{1}. If rzr_{z}^{*} is the adjoint operator, then rzλ𝒟r_{z}^{*}\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}. Let λn:𝒟\lambda_{n}:{\mathcal{D}}\to{\mathbb{C}} be given by

λn(Φ)=12πiS1zn1(Φ,rzλ)𝒟,𝒟𝑑z.\lambda_{n}(\Phi)=\frac{1}{2\pi{\mathrm{i}}}\int_{S^{1}}z^{-n-1}\left(\Phi,r_{z}^{*}\lambda\right)_{{\mathcal{D}},{\mathcal{D}}_{\mathcal{F}}^{*}}\,dz.

To see that the integral exists, observe that

(φ1(f1)φk(fk)Ω,rzλ)𝒟,𝒟=(φ1(βd1(rz)f1)φk(βdk(rz)fk)Ω,λ)𝒟,𝒟.\left(\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega,r_{z}^{*}\lambda\right)_{{\mathcal{D}},{\mathcal{D}}_{\mathcal{F}}^{*}}=\left(\varphi_{1}(\beta_{d_{1}}(r_{z})f_{1})\cdots\varphi_{k}(\beta_{d_{k}}(r_{z})f_{k})\Omega,\lambda\right)_{{\mathcal{D}},{\mathcal{D}}_{\mathcal{F}}^{*}}.

Since (z,f)βd(rz)f(z,f)\mapsto\beta_{d}(r_{z})f is jointly continuous S1×C(S1)C(S1)S^{1}\times C^{\infty}(S^{1})\to C^{\infty}(S^{1}), and λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}, the expression (φ1(f1)φk(fk)Ω,rzλ)\left(\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega,r_{z}^{*}\lambda\right) is jointly continuous S1×C(S1)kS^{1}\times C^{\infty}(S^{1})^{k}\to{\mathbb{C}}, and thus the integral defining λn\lambda_{n} exists. Moreover, we see that λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}. If u𝒱(m)u\in{\mathcal{V}}(m), then λn(u)=δn,mλ(u)\lambda_{n}(u)=\delta_{n,m}\lambda(u), so λn𝒟𝒱(n)\lambda_{n}\in{\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n)^{*} and λn(v)=λ(v)0\lambda_{n}(v)=\lambda(v)\neq 0.

For part (ii), it suffices to consider the {\mathcal{F}}-strong topology. Let 𝒲(n)𝒟{\mathcal{W}}(n)\subset{\mathcal{D}} be the subspace of vectors with conformal dimension nn. Note that

𝒱(n)=span{φ1(ej1)φk(ejk)Ω|ji=n}𝒲(n){\mathcal{V}}(n)=\operatorname{span}\{\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega\,|\,\sum j_{i}=-n\}\subset{\mathcal{W}}(n)

but equality is not immediate (it is e.g.​ a consequence of Proposition 3.14). Now fix n<0n<0 and let v𝒲(n)v\in{\mathcal{W}}(n). To verify the spectrum condition we must show that v=0v=0. Let λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}, and as above let λn=12πiS1zn1rzλ𝑑z\lambda_{n}=\frac{1}{2\pi{\mathrm{i}}}\int_{S^{1}}z^{-n-1}r_{z}^{*}\lambda\,dz. Then λn(v)=λ(v)\lambda_{n}(v)=\lambda(v), but λn\lambda_{n} vanishes on m0𝒱(m)\bigoplus_{m\geq 0}{\mathcal{V}}(m). Since the latter space is assumed to be dense, and λn𝒟\lambda_{n}\in{\mathcal{D}}_{\mathcal{F}}^{*} is continuous, we have λn0\lambda_{n}\equiv 0. Hence λ(v)=0\lambda(v)=0, and since 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} separates points we have v=0v=0, as desired. ∎

We will later see that for every λ𝒱(n)\lambda\in{\mathcal{V}}(n)^{*} there is a unique extension to a linear functional in 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} that vanishes on 𝒱(m){\mathcal{V}}(m) (when mnm\neq n) – see Proposition 3.14.

The notion of Wightman CFT presented in Definition 2.9 generalizes the unitary notion of Wightman CFT considered in [RTT22] in several ways. Most notably, the domain 𝒟{\mathcal{D}} does not have an inner product. We also do not assume that Ω\Omega is the unique Mo¨b\operatorname{M\ddot{o}b}-invariant vector up to scale, and we do not require that the eigenspaces for the generator of rotation are finite-dimensional.

The requirement that {\mathcal{F}} act regularly on 𝒟{\mathcal{D}} is necessary, as there exist pathological examples of quadruples (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) that satisfy all of the requirements to be a Wightman CFT except for the regularity of the action of {\mathcal{F}}. Indeed, we can refine Example 2.2 to produce a quadruple with non-regular action for which the only finite-energy vectors are scalar multiples of the vacuum, despite 𝒟{\mathcal{D}} being infinite-dimensional.

Example 2.12.

Let 𝒳<0C(S1){\mathcal{X}}_{<0}\subset C^{\infty}(S^{1}) be the closed span of z1,z2,z^{-1},z^{-2},\ldots. These are the functions ff in C(S1)C^{\infty}(S^{1}) that extend to holomorphic functions outside the unit disk which vanish at infinity. We similarly define 𝒳0{\mathcal{X}}_{\geq 0}. Let p:C(S1)𝒳<0p:C^{\infty}(S^{1})\to{\mathcal{X}}_{<0} be the projection with kernel 𝒳0{\mathcal{X}}_{\geq 0}. Let 𝒟=S(𝒳<0)=k=0Sk(𝒳<0){\mathcal{D}}=S({\mathcal{X}}_{<0})=\bigoplus_{k=0}^{\infty}S^{k}({\mathcal{X}}_{<0}) be the symmetric algebra, and let ΩS0(𝒳<0)\Omega\in S^{0}({\mathcal{X}}_{<0}) be the unit. Let φ\varphi be the operator-value distribution with domain 𝒟{\mathcal{D}} where φ(f)\varphi(f) acts by multiplication by pfpf in S(𝒳<0)S({\mathcal{X}}_{<0}). This action is evidently regular. The family ={φ}{\mathcal{F}}=\{\varphi\} acts locally on 𝒟{\mathcal{D}} since the symmetric algebra is abelian, and the vacuum vector is cyclic for {\mathcal{F}}.

Define a representation of Mo¨b\operatorname{M\ddot{o}b} on 𝒳<0{\mathcal{X}}_{<0} by U(γ)f=p(fγ1)U(\gamma)f=p(f\circ\gamma^{-1}), and extend this representation to S(𝒳<0)S({\mathcal{X}}_{<0}) (this representation is better understood as the quotient of the natural representation of Mo¨b\operatorname{M\ddot{o}b} on 𝒳0{\mathcal{X}}_{\leq 0} by constant functions). The representation has positive energy, and Ω\Omega is Mo¨b\operatorname{M\ddot{o}b}-invariant. Moreover we have for fC(S1)f\in C^{\infty}(S^{1})

U(γ)pf=p((pf)γ1)=p(fγ1)(as the constant component is annihilated by p)U(\gamma)pf=p((pf)\circ\gamma^{-1})=p(f\circ\gamma^{-1})\quad\text{(as the constant component is annihilated by }p\text{)}

and it follows that U(γ)φ(f)=φ(fγ1)U(γ)=φ(β1(γ)f)U(γ)U(\gamma)\varphi(f)=\varphi(f\circ\gamma^{-1})U(\gamma)=\varphi(\beta_{1}(\gamma)f)U(\gamma). Hence φ\varphi is Möbius covariant with conformal dimension 1, and we have shown that (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) is a Wightman CFT.

Now let 𝒳ωC(S1){\mathcal{X}}_{\omega}\subset C^{\infty}(S^{1}) be the dense subspace of functions which extend holomorphically to a neighborhood of S1S^{1}. Then 𝒳ω{\mathcal{X}}_{\omega} is invariant under U(γ)U(\gamma). Moreover, a function ff lies in 𝒳ω{\mathcal{X}}_{\omega} precisely when its Fourier coefficients decay sufficiently rapidly, and thus 𝒳ω{\mathcal{X}}_{\omega} is invariant under pp as well. Let 𝒟{\mathcal{I}}\subsetneq{\mathcal{D}} be the left ideal generated by p𝒳ω𝒳<0p{\mathcal{X}}_{\omega}\subset{\mathcal{X}}_{<0}, and observe that {\mathcal{I}} is invariant under φ(f)\varphi(f) and U(γ)U(\gamma) for all fC(S1)f\in C^{\infty}(S^{1}) and γMo¨b\gamma\in\operatorname{M\ddot{o}b}. Let 𝒟~=𝒟/\tilde{\mathcal{D}}={\mathcal{D}}/{\mathcal{I}}, let Ω~𝒟~\tilde{\Omega}\in\tilde{\mathcal{D}} be the image of Ω\Omega under the canonical projection, and let φ~(f)\tilde{\varphi}(f) and U~(γ)\tilde{U}(\gamma) be the operators on 𝒟~\tilde{\mathcal{D}} induced by φ(f)\varphi(f) and U(γ)U(\gamma), respectively. The quadruple (~,𝒟~,U~,Ω~)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega}) satisfy all of the requirements of a Wightman CFT except for regularity of the action of {\mathcal{F}}. We have

𝒱(n)=span{φ~1(ej1)φ~k(ejk)Ω~j1++jk=n}={0}{\mathcal{V}}(n)=\operatorname{span}\{\tilde{\varphi}_{1}(e_{j_{1}})\cdots\tilde{\varphi}_{k}(e_{j_{k}})\tilde{\Omega}\mid j_{1}+\cdots+j_{k}=-n\}=\{0\}

for all non-zero nn\in{\mathbb{Z}} since ej𝒳ωe_{j}\in{\mathcal{X}}_{\omega}.

Remark 2.13.

If (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) satisfy all of the requirements of a Wightman CFT except that the fields {\mathcal{F}} do not act regularly, then we can obtain a Wightman CFT on a quotient of 𝒟{\mathcal{D}}. Let 𝒟0=λ𝒟kerλ{\mathcal{D}}_{0}=\bigcap_{\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}}\operatorname{ker}\lambda. Since 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} is invariant under the adjoint actions φ(f)\varphi(f)^{*} and U(γ)U(\gamma)^{*}, it follows that 𝒟0{\mathcal{D}}_{0} is invariant under φ(f)\varphi(f) and U(γ)U(\gamma), and so we have actions of smeared fields and Mo¨b\operatorname{M\ddot{o}b} on the quotient 𝒟~=𝒟/𝒟0\tilde{\mathcal{D}}={\mathcal{D}}/{\mathcal{D}}_{0}. So long as 𝒟0𝒟{\mathcal{D}}_{0}\neq{\mathcal{D}}, these actions give a Wightman CFT on 𝒟~\tilde{\mathcal{D}}.

If one applies this procedure to the example (~,𝒟~,U~,Ω~)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega}) constructed in Example 2.12, then one obtains the trivial Wightman CFT. Indeed, we can check that 𝒟~=Ω~𝒟~0\tilde{\mathcal{D}}={\mathbb{C}}\tilde{\Omega}\oplus\tilde{\mathcal{D}}_{0} as follows. Returning to the notation of Example 2.12, we have =k=1Sk(𝒳<0){\mathcal{I}}=\bigoplus_{k=1}^{\infty}{\mathcal{I}}\cap S^{k}({\mathcal{X}}_{<0}), and thus

𝒟~=Ω~k=1Sk(𝒳<0)/(Sk(𝒳<0)).\tilde{\mathcal{D}}={\mathbb{C}}\tilde{\Omega}\oplus\bigoplus_{k=1}^{\infty}S^{k}({\mathcal{X}}_{<0})/({\mathcal{I}}\cap S^{k}({\mathcal{X}}_{<0})).

Hence it suffices to check that an arbitrary λ𝒟~~\lambda\in\tilde{\mathcal{D}}_{\tilde{\mathcal{F}}}^{*} vanishes on each of the spaces Sk(𝒳<0)/(Sk(𝒳<0))S^{k}({\mathcal{X}}_{<0})/({\mathcal{I}}\cap S^{k}({\mathcal{X}}_{<0})) when k1k\geq 1. Note that this space is spanned by vectors of the form φ~(f1)φ~(fk)Ω\tilde{\varphi}(f_{1})\cdots\tilde{\varphi}(f_{k})\Omega. When one of the functions fjf_{j} lies in 𝒳ω{\mathcal{X}}_{\omega} we have φ(f1)φ(fk)Ω\varphi(f_{1})\cdots\varphi(f_{k})\Omega\in{\mathcal{I}} and thus φ~(f1)φ~(fk)Ω=0\tilde{\varphi}(f_{1})\cdots\tilde{\varphi}(f_{k})\Omega=0. Hence for any λ𝒟~~\lambda\in\tilde{\mathcal{D}}_{\tilde{\mathcal{F}}}^{*} we have

λ(φ~(f1)φ~(fk)Ω)=0\lambda(\tilde{\varphi}(f_{1})\cdots\tilde{\varphi}(f_{k})\Omega)=0 (2.4)

when some fjf_{j} lies in 𝒳ω{\mathcal{X}}_{\omega}. Since 𝒳ω{\mathcal{X}}_{\omega} is dense in C(S1)C^{\infty}(S^{1}) and λ𝒟~~\lambda\in\tilde{\mathcal{D}}_{\tilde{\mathcal{F}}}^{*}, it follows that (2.4) holds for arbitrary functions fjC(S1)f_{j}\in C^{\infty}(S^{1}), and we conclude that λ\lambda vanishes on Sk(𝒳<0)/(Sk(𝒳<0))S^{k}({\mathcal{X}}_{<0})/({\mathcal{I}}\cap S^{k}({\mathcal{X}}_{<0})) when k1k\geq 1. Hence 𝒟~=Ω~𝒟~0\tilde{\mathcal{D}}={\mathbb{C}}\tilde{\Omega}\oplus\tilde{\mathcal{D}}_{0} as claimed, and 𝒟~~\tilde{\tilde{\mathcal{D}}} is spanned by the vacuum Ω~~\tilde{\tilde{\Omega}}.

With these distinctions in mind, we give the corresponding notion of vertex algebra after some brief preliminaries. Let Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}) be the three-dimensional real Lie algebra of Mo¨b=PSU(1,1)\operatorname{M\ddot{o}b}=\mathrm{PSU}(1,1). If Mo¨b\operatorname{M\ddot{o}b} is regarded as a subgroup of the group Diff(S1)\mathrm{Diff}(S^{1}) of orientation-preserving diffeomorphisms of the unit circle S1S^{1}, then Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}) is identified with a three-dimensional subspace of the space of smooth vector fields Vect(S1)\mathrm{Vect}(S^{1}) on S1S^{1}. Each vector field is identified with a differential operator f(eiϑ)ddϑf({\mathrm{e}}^{{\mathrm{i}}\vartheta})\frac{d}{d\vartheta} for some smooth function f(eiϑ)f({\mathrm{e}}^{{\mathrm{i}}\vartheta}), and the Lie bracket is given by [fddϑ,gddϑ]=(fgfg)ddϑ[f\frac{d}{d\vartheta},g\frac{d}{d\vartheta}]=(f^{\prime}g-fg^{\prime})\frac{d}{d\vartheta}, where ff^{\prime} denotes dfdϑ\tfrac{df}{d\vartheta}. Note that this bracket is the opposite of the bracket of vector fields, which is the natural choice when identifying Vect(S1)\mathrm{Vect}(S^{1}) with the Lie algebra of Diff(S1)\mathrm{Diff}(S^{1}). The complexification Lie(Mo¨b)𝔰𝔩(2,)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}}\cong\mathfrak{sl}(2,{\mathbb{C}}) of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}) is spanned by the elements {L1,L0,L1}\{L_{-1},L_{0},L_{1}\}, where LmL_{m} is the complexified vector field ieimϑddϑ-{\mathrm{i}}{\mathrm{e}}^{{\mathrm{i}}m\vartheta}\frac{d}{d\vartheta}. The vector fields LmL_{m} satisfy the commutation relations

[Lm,Ln]=(mn)Lm+n,m,n=1,0,1.\displaystyle[L_{m},L_{n}]=(m-n)L_{m+n},\qquad m,n=-1,0,1.

In a representation of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}}, we will frequently abuse notation and write LkL_{k} for the operator corresponding to the vector field indicated above.

If 𝒱{\mathcal{V}} is a vector space we write End(𝒱)[[z±1]]\operatorname{End}({\mathcal{V}})[[z^{\pm 1}]] for the vector space of formal power series in z±1z^{\pm 1} with coefficients in End(𝒱)\operatorname{End}({\mathcal{V}}). Given v𝒱v\in{\mathcal{V}} and A(z)=nAnznEnd(𝒱)[[z±1]]A(z)=\sum_{n\in{\mathbb{Z}}}A_{n}z^{n}\in\operatorname{End}({\mathcal{V}})[[z^{\pm 1}]], we have a formal series A(z)v=nAnvznA(z)v=\sum_{n}A_{n}vz^{n} with coefficients in 𝒱{\mathcal{V}}. For any BEnd(𝒱)B\in\operatorname{End}({\mathcal{V}}) we have [A(z),B]=n[An,B]znEnd(𝒱)[[z±1]][A(z),B]=\sum_{n}[A_{n},B]z^{n}\in\operatorname{End}({\mathcal{V}})[[z^{\pm 1}]]. If B(w)B(w) is another formal series in a second formal variable ww, then the expression [A(z),B(w)][A(z),B(w)] makes sense as a formal series in z±1z^{\pm 1} and w±1w^{\pm 1}.

We can now precisely specify the flavor of vertex algebras that we will consider.111The term Möbius vertex algebra has been used in the literature to describe various slightly different notions (see e.g.​ [BK08, HLZ14, Hua20, Kac98]). In some cases, authors include the possibility of fermionic fields, with the corresponding super version of the locality axiom; the term Möbius vertex superalgebra is also used in this case. Additionally, some authors replace our {\mathbb{N}}-grading with a more general grading. We do not foresee any significant obstacles to generalizing our results to Möbius vertex superalgebras graded by a lower-bounded subset of 12\tfrac{1}{2}{\mathbb{Z}}.

Definition 2.14.

An ({\mathbb{N}}-graded) Möbius vertex algebra consists of a vector space 𝒱{\mathcal{V}} equipped with a representation {L1,L0,L1}\{L_{-1},L_{0},L_{1}\} of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}}, a state-field correspondence Y:𝒱End(𝒱)[[z±1]]Y:{\mathcal{V}}\to\operatorname{End}({\mathcal{V}})[[z^{\pm 1}]], and a choice of non-zero vector Ω𝒱\Omega\in{\mathcal{V}} such that the following hold:

  1. (VA1)

    𝒱=n=0𝒱(n){\mathcal{V}}=\bigoplus_{n=0}^{\infty}{\mathcal{V}}(n), where 𝒱(n)=ker(L0n){\mathcal{V}}(n)=\ker(L_{0}-n).

  2. (VA2)

    Y(Ω,z)=Id𝒱Y(\Omega,z)=\operatorname{Id}_{\mathcal{V}} and Y(v,z)Ω|z=0=v\left.Y(v,z)\Omega\right|_{z=0}=v, i.e. Y(v,z)ΩY(v,z)\Omega has only non-negative powers of zz for all v𝒱v\in{\mathcal{V}}.

  3. (VA3)

    Ω\Omega is Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})-invariant, i.e.​ LmΩ=0L_{m}\Omega=0 for m=1,0,1m=-1,0,1.

  4. (VA4)

    [Lm,Y(v,z)]=j=0m+1(m+1j)zm+1jY(Lj1v,z)[L_{m},Y(v,z)]=\sum_{j=0}^{m+1}\binom{m+1}{j}z^{m+1-j}Y(L_{j-1}v,z) and Y(L1v,z)=ddzY(v,z)Y(L_{-1}v,z)=\frac{d}{dz}Y(v,z) for all v𝒱v\in{\mathcal{V}} and m=1,0,1m=-1,0,1.

  5. (VA5)

    (zw)N[Y(v,z),Y(u,w)]=0(z-w)^{N}[Y(v,z),Y(u,w)]=0 for NN sufficiently large.

For v𝒱v\in{\mathcal{V}}, we write Y(v,z)=mv(m)zm1Y(v,z)=\sum_{m\in{\mathbb{Z}}}v_{(m)}z^{-m-1}, where v(m)End(𝒱)v_{(m)}\in\operatorname{End}({\mathcal{V}}) are called the modes of vv. A vector v𝒱v\in{\mathcal{V}} is called homogeneous (with conformal dimension dd) if it lies in 𝒱(d){\mathcal{V}}(d). As a consequence of the L0L_{0}-commutation relation, when vv is homogeneous with conformal dimension dd we have [L0,v(m)]=(dm1)v(m)[L_{0},v_{(m)}]=(d-m-1)v_{(m)}. Hence v(m)v_{(m)} maps 𝒱(n){\mathcal{V}}(n) into 𝒱(n+dm1){\mathcal{V}}(n+d-m-1). A vector v𝒱v\in{\mathcal{V}} is called quasiprimary if it is homogeneous and L1v=0L_{1}v=0.

We record two useful identities satisfied by the modes of a vertex operator (see [Kac98, §4.8]), the Borcherds product formula:

(u(n)v)(k)=j=0(1)j(nj)(u(nj)v(k+j)(1)nv(n+kj)u(j)),\big{(}u_{(n)}v\big{)}_{(k)}=\sum_{j=0}^{\infty}(-1)^{j}\binom{n}{j}\left(u_{(n-j)}v_{(k+j)}-(-1)^{n}v_{(n+k-j)}u_{(j)}\right), (2.5)

and the Borcherds commutator formula:

[u(m),v(k)]=j=0(mj)(u(j)v)(m+kj).[u_{(m)},v_{(k)}]=\sum_{j=0}^{\infty}\binom{m}{j}\big{(}u_{(j)}v\big{)}_{(m+k-j)}. (2.6)

Note that when the sums of operators on the right-hand sides of (2.5) and (2.6) are applied to a vector, all but finitely many terms vanish.

3 Equivalence between Möbius vertex algebras and Wightman CFTs

3.1 From vertex algebras to Wightman CFTs

In this section we construct a Wightman CFT from a Möbius vertex algebra 𝒱=n=0𝒱(n){\mathcal{V}}=\bigoplus_{n=0}^{\infty}{\mathcal{V}}(n). The first step is to construct operator-valued distributions from the formal distributions Y(v,z)Y(v,z), as follows. For v𝒱(d)v\in{\mathcal{V}}(d), the degree-shifted mode vnv_{n} is defined by vn:=v(n+d1)v_{n}:=v_{(n+d-1)}, which gives an alternative field expansion Y(v,z)=n=vnzndY(v,z)=\sum_{n=-\infty}^{\infty}v_{n}z^{-n-d}, so that vn𝒱(m)𝒱(mn)v_{n}{\mathcal{V}}(m)\subset{\mathcal{V}}(m-n). We extend the definition of vnv_{n} to non-homogeneous vectors by linearity. Let us write 𝒱{\mathcal{V}}^{\prime} for the restricted dual 𝒱=n=0𝒱(n){\mathcal{V}}^{\prime}=\bigoplus_{n=0}^{\infty}{\mathcal{V}}(n)^{*}, which is to say linear functionals on 𝒱{\mathcal{V}} that are supported on finitely many 𝒱(n){\mathcal{V}}(n). We denote by 𝒱^\widehat{\mathcal{V}} the algebraic completion

𝒱^=n=0𝒱(n),\widehat{\mathcal{V}}=\prod_{n=0}^{\infty}{\mathcal{V}}(n),

and we embed 𝒱𝒱^{\mathcal{V}}\subset\widehat{\mathcal{V}} in the natural way. We equip 𝒱^\widehat{\mathcal{V}} with the weak topology induced by the pairing with 𝒱{\mathcal{V}}^{\prime}.

For fC(S1)f\in C^{\infty}(S^{1}) we define

Y0(v,f):𝒱𝒱^Y^{0}(v,f):{\mathcal{V}}\to\widehat{\mathcal{V}}

by

Y0(v,f)u=nf^(n)vnu,Y^{0}(v,f)u=\sum_{n\in{\mathbb{Z}}}\hat{f}(n)v_{n}u,

where f^(n)\hat{f}(n) is the nn-th Fourier coefficient of ff.

We now show that the maps Y0(v,f)Y^{0}(v,f) may be extended to act on an invariant domain 𝒟𝒱^{\mathcal{D}}\subset\widehat{\mathcal{V}}. The first step is the following lemma, which has an identical proof to [RTT22, Lem.​ 2.7].

Lemma 3.1.

For all v1,,vk,u𝒱v^{1},\ldots,v^{k},u\in{\mathcal{V}} and u𝒱u^{\prime}\in{\mathcal{V}}^{\prime}, there exists a polynomial pp such that

|(vm11vm22vmkku,u)𝒱,𝒱||p(m1,,mk)|\big{|}\left(v^{1}_{m_{1}}v^{2}_{m_{2}}\cdots v^{k}_{m_{k}}u,u^{\prime}\right)_{{\mathcal{V}},{\mathcal{V}}^{\prime}}\big{|}\leq\left|p(m_{1},\ldots,m_{k})\right|

for all (m1,,mk)k(m_{1},\ldots,m_{k})\in{\mathbb{Z}}^{k}. The polynomial depends on the vectors vj,u,v^{j},u, and uu^{\prime}, but the degree of pp may be bounded independent of uu and uu^{\prime}.

We refer the reader to [RTT22] for a proof. A very similar argument, however, yields the following observation.

Lemma 3.2.

Let v1,,vk𝒱v^{1},\ldots,v^{k}\in{\mathcal{V}} and let n0n\in{\mathbb{Z}}_{\geq 0}. Then

𝒱(n;v1,,vk):=span{vm11vmkkΩ|m1++mk=n}{\mathcal{V}}(n;v^{1},\cdots,v^{k}):=\operatorname{span}\{v^{1}_{m_{1}}\cdots v_{m_{k}}^{k}\Omega\,|\,m_{1}+\cdots+m_{k}=-n\}

is finite-dimensional.

Proof.

We proceed by induction on kk, with the cases k=0k=0 and k=1k=1 being immediate. Now fix k2k\geq 2, and suppose dim𝒱(n;u1,,u)<\dim{\mathcal{V}}(n^{\prime};u^{1},\cdots,u^{\ell})<\infty when <k\ell<k. First observe that when m1<nm_{1}<-n and m1+mk=nm_{1}+\cdots m_{k}=-n we must have m2+mk>0m_{2}+\cdots m_{k}>0, and hence

vm11vmkkΩ=0.v^{1}_{m_{1}}\cdots v_{m_{k}}^{k}\Omega=0.

Next observe that

𝒱(n;v1,,vk)=\displaystyle{\mathcal{V}}(n;v^{1},\cdots,v^{k})=
m1=n0vm11𝒱(n+m1;v2,,vk)+span{vm11vmkkΩ|m1++mk=n,m1>0}\displaystyle\quad\sum_{{m_{1}}=-n}^{0}v^{1}_{m_{1}}{\mathcal{V}}(n+{m_{1}};v^{2},\ldots,v^{k})+\operatorname{span}\{v^{1}_{m_{1}}\cdots v_{m_{k}}^{k}\Omega\,|\,m_{1}+\cdots+m_{k}=-n,\,m_{1}>0\}

and each of the subspaces vm11𝒱(n+m1;v2,,vk)v^{1}_{m_{1}}{\mathcal{V}}(n+m_{1};v^{2},\ldots,v^{k}) is finite-dimensional by the inductive hypothesis, so it suffices to show that the final term is finite-dimensional as well.

By the Borcherds commutator formula (2.6), if m1>0m_{1}>0 and m1+mk=nm_{1}+\cdots m_{k}=-n we have

vm11vmkkΩ=j=2kvm22[vm11,vmjj]vmkkΩj=2ks=0d1+dj1𝒱(n;v2,,vsd1+11vj,vk).v^{1}_{m_{1}}\cdots v_{m_{k}}^{k}\Omega=\sum_{j=2}^{k}v^{2}_{m_{2}}\cdots[v^{1}_{m_{1}},v^{j}_{m_{j}}]\cdots v^{k}_{m_{k}}\Omega\in\sum_{j=2}^{k}\sum_{s=0}^{d_{1}+d_{j}-1}{\mathcal{V}}(n;v^{2},\ldots,v^{1}_{s-d_{1}+1}v^{j},\ldots v^{k}).

The subspace on the right-hand side is finite-dimensional by the inductive hypothesis and independent of m1,,mkm_{1},\ldots,m_{k}, and thus we conclude that 𝒱(n;v1,,vk){\mathcal{V}}(n;v^{1},\ldots,v^{k}) is finite-dimensional as well. ∎

If f[z±1]f\in{\mathbb{C}}[z^{\pm 1}], then Y0(v,f)Y^{0}(v,f) maps 𝒱{\mathcal{V}} into 𝒱{\mathcal{V}}. Our next lemma gives an estimate for these maps Y0(v,f)Y^{0}(v,f) in terms of the NN-Sobolev norm of ff. Recall that for N0N\in{\mathbb{R}}_{\geq 0}, the NN-Sobolev norm on C(S1)C^{\infty}(S^{1}) is given by

fN=(n|f^(n)|2(1+n2)N)1/2.\left\|f\right\|_{N}=\left(\sum_{n\in{\mathbb{Z}}}\big{|}\hat{f}(n)\big{|}^{2}(1+n^{2})^{N}\right)^{1/2}. (3.1)

We denote by HN(S1)H^{N}(S^{1}) the Hilbert space completion of C(S1)C^{\infty}(S^{1}) under this norm, which consists of L2L^{2}-functions with finite NN-Sobolev norm. The locally convex topology on C(S1)C^{\infty}(S^{1}) is induced by the norms N\left\|\cdot\right\|_{N}, and a linear map from C(S1)C^{\infty}(S^{1}) to a Banach space is continuous precisely when it is bounded with respect to some NN-Sobolev norm. We then have the following estimate, which is a simplification of [RTT22, Lem.​ 2.8].

Lemma 3.3.

For all v1,,vk,u𝒱v_{1},\ldots,v_{k},u\in{\mathcal{V}}, u𝒱u^{\prime}\in{\mathcal{V}}^{\prime}, and Laurent polynomials f1,,fk[z±1]f_{1},\ldots,f_{k}\in{\mathbb{C}}[z^{\pm 1}], we have

|(Y0(vk,fk)Y0(v1,f1)u,u)𝒱,𝒱|Cf1NfkN.\left|\left(Y^{0}(v_{k},f_{k})\cdots Y^{0}(v_{1},f_{1})u,u^{\prime}\right)_{{\mathcal{V}},{\mathcal{V}}^{\prime}}\right|\leq C\left\|f_{1}\right\|_{N}\cdots\left\|f_{k}\right\|_{N}.

The number NN depends only on the vjv_{j}, and the constant CC depends on the {vj}j=1,,k\{v_{j}\}_{j=1,\cdots,k}, uu, and uu^{\prime}.

The auxiliary domain 𝒟{\mathcal{D}} and the topology on it.

By Lemma 3.2, the assignment (f1,,fk)Y0(vk,fk)Y0(v1,f1)u(f_{1},\ldots,f_{k})\mapsto Y^{0}(v_{k},f_{k})\cdots Y^{0}(v_{1},f_{1})u gives a map

[z±1]kn=0𝒱(n;v1,,vk,u){\mathbb{C}}[z^{\pm 1}]^{k}\to\prod_{n=0}^{\infty}{\mathcal{V}}(n;v_{1},\ldots,v_{k},u)

with each space 𝒱(n;v1,,vk,u){\mathcal{V}}(n;v_{1},\ldots,v_{k},u) finite-dimensional. By Lemma 3.3, this extends to a continuous multilinear map again taking values in n=0𝒱(n;v1,,vk,u)𝒱^\prod_{n=0}^{\infty}{\mathcal{V}}(n;v_{1},\ldots,v_{k},u)\subset\widehat{\mathcal{V}}. Thus for each v1,,vk,u𝒱v_{1},\ldots,v_{k},u\in{\mathcal{V}}, there exists a unique continuous multilinear map

Xv1,,vk,u:C(S1)k𝒱^X_{v_{1},\ldots,v_{k},u}:C^{\infty}(S^{1})^{k}\to\widehat{\mathcal{V}}

such that when f1,,fk[z±1]f_{1},\ldots,f_{k}\in{\mathbb{C}}[z^{\pm 1}] we have

Xv1,,vk,u(f1,,fk)=Y0(vk,fk)Y0(v1,f1)u.X_{v_{1},\ldots,v_{k},u}(f_{1},\ldots,f_{k})=Y^{0}(v_{k},f_{k})\cdots Y^{0}(v_{1},f_{1})u. (3.2)

Let 𝒟0=Ω{\mathcal{D}}_{0}={\mathbb{C}}\Omega, and for k=1,2,k=1,2,\ldots set

𝒟k=span{Xv1,,vk,Ω(f1,,fk)|vj𝒱,fjC(S1)}𝒱^.{\mathcal{D}}_{k}=\operatorname{span}\{X_{v_{1},\ldots,v_{k},\Omega}(f_{1},\ldots,f_{k})\,|\,v_{j}\in{\mathcal{V}},f_{j}\in C^{\infty}(S^{1})\}\subset\widehat{\mathcal{V}}.

We have 𝒟k𝒟k+1{\mathcal{D}}_{k}\subset{\mathcal{D}}_{k+1} by considering v1=Ωv_{1}=\Omega. Let 𝒟=k=0𝒟k𝒱^{\mathcal{D}}=\bigcup_{k=0}^{\infty}{\mathcal{D}}_{k}\subset\widehat{\mathcal{V}}, equipped with the subspace topology (i.e.​ the weak topology induced by the linear functionals 𝒱{\mathcal{V}}^{\prime}, in which a sequence (or net) Φj𝒟\Phi_{j}\in{\mathcal{D}} converges to Φ\Phi if and only if λ(Φj)\lambda(\Phi_{j}) converges to λ(Φ)\lambda(\Phi) for all λ𝒱\lambda\in{\mathcal{V}}^{\prime}).

Lemma 3.4.

For all v𝒱v\in{\mathcal{V}} and fC(S1)f\in C^{\infty}(S^{1}) there exists a unique continuous linear map Y(v,f):𝒟𝒟Y(v,f):{\mathcal{D}}\to{\mathcal{D}} such that:

  1. i)

    Y(v,f)|𝒱=Y0(v,f)Y(v,f)|_{{\mathcal{V}}}=Y^{0}(v,f).

  2. ii)

    The expressions Y(v1,f1)Y(vk,fk)ΩY(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega are (jointly) continuous in the functions fjf_{j}.

In addition, we have 𝒟=span{Y(v1,f1)Y(vk,fk)Ω|k0,vj𝒱,fjC(S1)}{\mathcal{D}}=\operatorname{span}\{Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega\,|\,k\in{\mathbb{Z}}_{\geq 0},v_{j}\in{\mathcal{V}},f_{j}\in C^{\infty}(S^{1})\}.

Proof.

We first consider uniqueness. When f1,,fkf_{1},\ldots,f_{k} are Laurent polynomials, the condition Y(v,f)|𝒱=Y0(v,f)Y(v,f)|_{{\mathcal{V}}}=Y^{0}(v,f) determines the value of Y(v1,f1)Y(vk,fk)Ω𝒱Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega\in{\mathcal{V}}. The value of such expressions in 𝒟{\mathcal{D}} is then uniquely determined by continuity in the functions fjf_{j}.

We now show existence. We wish to define Y(v,f)Y(v,f) on Xv1,,vk,Ω(f1,,fk)𝒟kX_{v_{1},\ldots,v_{k},\Omega}(f_{1},\ldots,f_{k})\in{\mathcal{D}}_{k} by the formula

Y(v,f)Xv1,,vk,Ω(f1,,fk)=Xv,v1,,vk,Ω(f,f1,,fk),Y(v,f)X_{v_{1},\ldots,v_{k},\Omega}(f_{1},\ldots,f_{k})=X_{v,v_{1},\ldots,v_{k},\Omega}(f,f_{1},\ldots,f_{k}), (3.3)

but must check that this is well-defined.

First, consider if ff is a Laurent polynomial. The modes vnv_{n} map 𝒱(m){\mathcal{V}}(m) to 𝒱(mn){\mathcal{V}}(m-n), and thus the adjoint (transpose) operator vnv_{n}^{*} maps 𝒱(m){\mathcal{V}}(m)^{*} into 𝒱(m+n){\mathcal{V}}(m+n)^{*}. Hence there is an adjoint map Y0(v,f):𝒱𝒱Y^{0}(v,f)^{*}:{\mathcal{V}}^{\prime}\to{\mathcal{V}}^{\prime} such that for u𝒱u\in{\mathcal{V}} and u𝒱u^{\prime}\in{\mathcal{V}}^{\prime} we have

(Y0(v,f)u,u)𝒱,𝒱=(u,Y0(v,f)u)𝒱,𝒱.\left(Y^{0}(v,f)u,u^{\prime}\right)_{{\mathcal{V}},{\mathcal{V}}^{\prime}}=\left(u,Y^{0}(v,f)^{*}u^{\prime}\right)_{{\mathcal{V}},{\mathcal{V}}^{\prime}}.

Thus if f,f1,,fkf,f_{1},\ldots,f_{k} are Laurent polynomials we have

(Xv,v1,,vk,u(f,f1,,fk),u)\displaystyle\left(X_{v,v_{1},\ldots,v_{k},u}(f,f_{1},\ldots,f_{k}),u^{\prime}\right) =(Y0(v,f)Y0(vk,fk)Y0(v1,f1)u,u)\displaystyle=\left(Y^{0}(v,f)Y^{0}(v_{k},f_{k})\cdots Y^{0}(v_{1},f_{1})u,u^{\prime}\right)
=(Y0(vk,fk)Y0(v1,f1)u,Y0(v,f)u)\displaystyle=\left(Y^{0}(v_{k},f_{k})\cdots Y^{0}(v_{1},f_{1})u,Y^{0}(v,f)^{*}u^{\prime}\right)
=(Xv1,,vk,u(f1,,fk),Y0(v,f)u).\displaystyle=\left(X_{v_{1},\ldots,v_{k},u}(f_{1},\ldots,f_{k}),Y^{0}(v,f)^{*}u^{\prime}\right).

As the first and last terms are jointly continuous in f1,,fkf_{1},\ldots,f_{k} by (3.2), we have

(Xv,v1,,vk,u(f,f1,,fk),u)𝒱^,𝒱=(Xv1,,vk,u(f1,,fk),Y0(v,f)u)𝒱^,𝒱\left(X_{v,v_{1},\ldots,v_{k},u}(f,f_{1},\ldots,f_{k}),u^{\prime}\right)_{\widehat{\mathcal{V}},{\mathcal{V}}^{\prime}}=\left(X_{v_{1},\ldots,v_{k},u}(f_{1},\ldots,f_{k}),Y^{0}(v,f)^{*}u^{\prime}\right)_{\widehat{\mathcal{V}},{\mathcal{V}}^{\prime}} (3.4)

whenever ff is a Laurent polynomial.

We now argue that (3.3) is well-defined. Let X~k:(𝒱C(S1))k𝒟\tilde{X}_{k}:({\mathcal{V}}\otimes C^{\infty}(S^{1}))^{\otimes k}\to{\mathcal{D}} be the linear map corresponding to the multilinear map Xv1,,vk,Ω(f1,,fk)X_{v_{1},\ldots,v_{k},\Omega}(f_{1},\ldots,f_{k}), so that 𝒟k{\mathcal{D}}_{k} is the range of X~k\tilde{X}_{k}. Let 𝒯=k=0(𝒱C(S1))k{\mathcal{T}}=\bigoplus_{k=0}^{\infty}({\mathcal{V}}\otimes C^{\infty}(S^{1}))^{\otimes k}, and let X~:𝒯𝒟\tilde{X}:{\mathcal{T}}\to{\mathcal{D}} be the map given by X~k\tilde{X}_{k} on the kkth direct summand of 𝒯{\mathcal{T}}. We wish to show that if Ξ𝒯\Xi\in{\mathcal{T}} and X~(Ξ)=0\tilde{X}(\Xi)=0, then X~(vfΞ)=0\tilde{X}(v\otimes f\otimes\Xi)=0 as well.

Fix Ξ\Xi as above. By (3.4), if ff is a Laurent polynomial we have for all u𝒱u^{\prime}\in{\mathcal{V}}^{\prime}

(X~(vfΞ),u)=(X~(Ξ),Y0(v,f)u)=0,\left(\tilde{X}(v\otimes f\otimes\Xi),u^{\prime}\right)=\left(\tilde{X}(\Xi),Y^{0}(v,f)^{*}u^{\prime}\right)=0,

and so X~(vfΞ)=0\tilde{X}(v\otimes f\otimes\Xi)=0. On the other hand, X~(vfΞ)\tilde{X}(v\otimes f\otimes\Xi) is continuous in ff, and so X~(vfΞ)=0\tilde{X}(v\otimes f\otimes\Xi)=0 vanishes for all fC(S1)f\in C^{\infty}(S^{1}). Thus there is a well-defined map Y(v,f):𝒟𝒟Y(v,f):{\mathcal{D}}\to{\mathcal{D}} satisfying (3.3).

By construction we have Y(v1,f1)Y(vk,fk)Ω=Xv1,,vk,Ω(f1,,fk)Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega=X_{v_{1},\ldots,v_{k},\Omega}(f_{1},\ldots,f_{k}). It follows immediately that such expressions span 𝒟{\mathcal{D}}, and since Xv1,,vk,ΩX_{v_{1},\ldots,v_{k},\Omega} is continuous in the functions fjf_{j} we have also shown the second required property of the operator Y(v,f)Y(v,f). For the first property, note that Y(v,f)|𝒱Y(v,f)|_{{\mathcal{V}}} agrees with Y0(v,f)Y^{0}(v,f) by (3.2) when ff is a Laurent polynomial, and thus for all ff by continuity. ∎

Remark 3.5.

From Lemma 3.3 and Lemma 3.4 we have shown a version of the uniformly bounded order property for the operator-valued distributions Y(v,f)Y(v,f), namely that for any v1,,vk𝒱v_{1},\ldots,v_{k}\in{\mathcal{V}}, there is a positive number NN such that for every u𝒱u\in{\mathcal{V}} the map (f1,,fk)Y(v1,f1)Y(vk,fk)u(f_{1},\ldots,f_{k})\mapsto Y(v_{1},f_{1})\ldots Y(v_{k},f_{k})u extends to a continuous map HN(S1)k𝒱^H^{N}(S^{1})^{k}\to\widehat{\mathcal{V}}.

We have constructed a family of operator-valued distributions Y(v,f)Y(v,f) on 𝒟{\mathcal{D}}. We next consider Möbius covariance of these distributions, which will hold when vv is quasiprimary. To this end we introduce

𝒱QP=span{v𝒱|v is quasiprimary}=kerL1.{\mathcal{V}}_{\mathrm{QP}}=\operatorname{span}\{v\in{\mathcal{V}}\,|\,v\text{ is quasiprimary}\}=\ker L_{1}.

Note that 𝒱QP=n=0𝒱QP𝒱(n){\mathcal{V}}_{\mathrm{QP}}=\bigoplus_{n=0}^{\infty}{\mathcal{V}}_{\mathrm{QP}}\cap{\mathcal{V}}(n), so that every vector in 𝒱QP{\mathcal{V}}_{\mathrm{QP}} may be written uniquely as a sum of homogeneous quasiprimary components. Let

𝒟QP=span{Y(v1,f1)Y(vk,fk)Ω|k0,vj𝒱QP,fjC(S1)}𝒟.{\mathcal{D}}_{\mathrm{QP}}=\operatorname{span}\{Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega\,|\,k\in{\mathbb{Z}}_{\geq 0},v_{j}\in{\mathcal{V}}_{\mathrm{QP}},f_{j}\in C^{\infty}(S^{1})\}\subseteq{\mathcal{D}}.

Let us assume that 𝒱{\mathcal{V}} is generated by 𝒱QP{\mathcal{V}}_{\mathrm{QP}} as a vertex algebra, in which case 𝒱𝒟QP{\mathcal{V}}\subset{\mathcal{D}}_{\mathrm{QP}}. If 𝒱{\mathcal{V}} is completely reducible as a Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})-module (i.e.​ if it is spanned by vectors of the form L1kvL_{-1}^{k}v with v𝒱QPv\in{\mathcal{V}}_{\mathrm{QP}}), then it is evidently generated by 𝒱QP{\mathcal{V}}_{\mathrm{QP}} as a vertex algebra and moreover 𝒟QP=𝒟{\mathcal{D}}_{\mathrm{QP}}={\mathcal{D}} as Y(L1v,f)=Y(v,ifmf)Y(L_{-1}v,f)=Y(v,{\mathrm{i}}f^{\prime}-mf) when v𝒱(m)v\in{\mathcal{V}}(m).

We will now construct a representation of U:Mo¨b(𝒟QP)U:\operatorname{M\ddot{o}b}\to{\mathcal{L}}({\mathcal{D}}_{\mathrm{QP}}) for which the Wightman fields Y(v,f)|𝒟QPY(v,f)|_{{\mathcal{D}}_{\mathrm{QP}}} are covariant for all v𝒱QPv\in{\mathcal{V}}_{\mathrm{QP}} and for which the vacuum Ω\Omega is invariant. Note that such a representation is unique if it exists, as the covariance condition implies

U(γ)Y(v1,f1)Y(vk,fk)Ω=Y(v1,βd1(γ)f1)Y(vk,βdk(γ)fk)Ω.U(\gamma)Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega=Y(v_{1},\beta_{d_{1}}(\gamma)f_{1})\cdots Y(v_{k},\beta_{d_{k}}(\gamma)f_{k})\Omega.

So the difficulty is in showing that a linear map satisfying the above condition exists, i.e.​ showing that if a linear combination of vectors of the form

Y(v1,f1)Y(vk,fk)ΩY(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega

vanishes, then so does the corresponding linear combination of

Y(v1,βd1(γ)f1)Y(vk,βdk(γ)fk)Ω.Y(v_{1},\beta_{d_{1}}(\gamma)f_{1})\cdots Y(v_{k},\beta_{d_{k}}(\gamma)f_{k})\Omega.

We first extend the representation of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}} furnished by the Möbius vertex algebra structure on 𝒱{\mathcal{V}} to a representation on 𝒟QP{\mathcal{D}}_{\mathrm{QP}}. Recall that Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}} is spanned by complexified vector fields g(eiϑ)ddϑg({\mathrm{e}}^{{\mathrm{i}}\vartheta})\frac{d}{d\vartheta} on the circle, where LkL_{k} corresponds to ieikϑddϑ-{\mathrm{i}}{\mathrm{e}}^{{\mathrm{i}}k\vartheta}\frac{d}{d\vartheta}.

Let 𝒱{\mathcal{V}}^{*} be the algebraic dual of 𝒱{\mathcal{V}}, and note that the adjoint operators Lk:𝒱𝒱L_{k}^{*}:{\mathcal{V}}^{*}\to{\mathcal{V}}^{*} leave 𝒱{\mathcal{V}}^{\prime} invariant. We claim that the closure of the graph Γ(Lk)𝒱×𝒱\Gamma(L_{k})\subset{\mathcal{V}}\times{\mathcal{V}} in 𝒱^×𝒱^\widehat{\mathcal{V}}\times\widehat{\mathcal{V}} is the graph of a densely defined linear operator. Indeed, suppose that vjv_{j} is a net in 𝒱{\mathcal{V}} such that vj0v_{j}\to 0 and LkvjvL_{k}v_{j}\to v in 𝒱^\widehat{\mathcal{V}}. Then for any λ𝒱\lambda\in{\mathcal{V}}^{\prime} it holds that

λ(v)=limjλ(Lkvj)=limj(Lkλ)(vj)=0.\lambda(v)=\lim_{j}\lambda(L_{k}v_{j})=\lim_{j}(L_{k}^{*}\lambda)(v_{j})=0.

As 𝒱{\mathcal{V}}^{\prime} separates points in 𝒱^\widehat{\mathcal{V}}, we conclude that v=0v=0 and that the closure of Γ(Lk)\Gamma(L_{k}) is the graph of a densely defined operator as claimed. Taking linear combinations we obtain a densely-defined operator on 𝒱^\widehat{\mathcal{V}} for every XLie(Mo¨b)X\in\operatorname{Lie}(\operatorname{M\ddot{o}b}), which we denote by π(X)\pi(X).

Lemma 3.6.

Let 𝒱{\mathcal{V}} be a Möbius vertex algebra that is generated as a vertex algebra by its quasiprimary fields. Then for any gddϑLie(Mo¨b)g\tfrac{d}{d\vartheta}\in\operatorname{Lie}(\operatorname{M\ddot{o}b}) the domain of π(gddϑ)\pi(g\tfrac{d}{d\vartheta}) contains 𝒟QP{\mathcal{D}}_{\mathrm{QP}} and π(gddϑ)\pi(g\tfrac{d}{d\vartheta}) leaves 𝒟QP{\mathcal{D}}_{\mathrm{QP}} invariant. Moreover, if vv is quasiprimary with conformal dimension dd we have

[π(gddϑ),Y(v,f)]=Y(v,(d1)dgdϑfgdfdϑ)[\pi(g\tfrac{d}{d\vartheta}),Y(v,f)]=Y(v,(d-1)\tfrac{dg}{d\vartheta}f-g\tfrac{df}{d\vartheta})

as endomorphisms of 𝒟QP{\mathcal{D}}_{\mathrm{QP}}.

Proof.

If v𝒱QP(d)v\in{\mathcal{V}}_{\mathrm{QP}}(d) then the commutation relations between Y(v,z)Y(v,z) and the LkL_{k} from the definition of a Möbius vertex algebra imply that when f[z±1]f\in{\mathbb{C}}[z^{\pm 1}] is a Laurent polynomial we have

[π(gddϑ),Y(v,f)]=Y(v,(d1)dgdϑfgdfdϑ)[\pi(g\tfrac{d}{d\vartheta}),Y(v,f)]=Y\big{(}v,(d-1)\tfrac{dg}{d\vartheta}f-g\tfrac{df}{d\vartheta}\big{)}

as endomorphisms of 𝒱{\mathcal{V}}. Thus if f1,,fkf_{1},\ldots,f_{k} are Laurent polynomials and v1,,vk𝒱QPv_{1},\ldots,v_{k}\in{\mathcal{V}}_{\mathrm{QP}}, then we have

π(gddϑ)Y(v1,f1)Y(vk,fk)Ω=j=1kY(v1,f1)Y(vj,(dj1)dgdϑfjgdfjdϑ)Y(vk,fk)Ω\displaystyle\pi(g\tfrac{d}{d\vartheta})Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega=\sum_{j=1}^{k}Y(v_{1},f_{1})\cdots Y(v_{j},(d_{j}-1)\tfrac{dg}{d\vartheta}f_{j}-g\tfrac{df_{j}}{d\vartheta})\cdots Y(v_{k},f_{k})\Omega (3.5)

where djd_{j} is the conformal dimension of vjv_{j}. For arbitrary f1,,fkC(S1)f_{1},\ldots,f_{k}\in C^{\infty}(S^{1}), choose sequences of Laurent polynomials222fj,nC(S1)f_{j,n}\in C^{\infty}(S^{1}) and it is not the nn-th Fourier coefficient of fjf_{j}. fj,nf_{j,n} such that limnfj,n=fj\lim_{n\to\infty}f_{j,n}=f_{j} in C(S1)C^{\infty}(S^{1}), and observe that

limnY(v1,f1,n)Y(vk,fk,n)Ω=Y(v1,f1)Y(vk,fk)Ω\lim_{n\to\infty}Y(v_{1},f_{1,n})\cdots Y(v_{k},f_{k,n})\Omega=Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega

and

limnY(v1,f1,n)Y(vj,(dj1)dgdϑfj,ngdfj,ndϑ)Y(vk,fk,n)Ω=\displaystyle\lim_{n\to\infty}Y(v_{1},f_{1,n})\cdots Y(v_{j},(d_{j}-1)\tfrac{dg}{d\vartheta}f_{j,n}-g\tfrac{df_{j,n}}{d\vartheta})\cdots Y(v_{k},f_{k,n})\Omega=
Y(v1,f1)Y(vj,(dj1)dgdϑfjgdfjdϑ)Y(vk,fk)Ω\displaystyle\quad Y(v_{1},f_{1})\cdots Y(v_{j},(d_{j}-1)\tfrac{dg}{d\vartheta}f_{j}-g\tfrac{df_{j}}{d\vartheta})\cdots Y(v_{k},f_{k})\Omega

in 𝒱^\widehat{\mathcal{V}} by Lemma 3.4. Hence Y(v1,f1)Y(vk,fk)ΩY(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega lies in the domain of π(gddϑ)\pi(g\tfrac{d}{d\vartheta}) and (3.5) holds for fjC(S1)f_{j}\in C^{\infty}(S^{1}). It follows that π(gddϑ)\pi(g\tfrac{d}{d\vartheta}) leaves 𝒟QP{\mathcal{D}}_{\mathrm{QP}} invariant and we have the desired commutation relation with smeared fields. ∎

We now turn to constructing the desired representation U:Mo¨b(𝒟QP)U:\operatorname{M\ddot{o}b}\to{\mathcal{L}}({\mathcal{D}}_{\mathrm{QP}}). The following lemma will allow us to define U(γ)U(\gamma) on a 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}}.

Lemma 3.7.

Let 𝒱{\mathcal{V}} be a Möbius vertex algebra which is generated by a set of quasiprimary vectors. Then for any γMo¨b\gamma\in\operatorname{M\ddot{o}b} there exists a unique linear map U0(γ):𝒱𝒟QPU^{0}(\gamma):{\mathcal{V}}\to{\mathcal{D}}_{\mathrm{QP}} such that

U0(γ)Y(v1,f1)Y(vk,fk)Ω=Y(v1,βd1(γ)f1)Y(vk,βdk(γ)fk)Ω.U^{0}(\gamma)Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega=Y(v_{1},\beta_{d_{1}}(\gamma)f_{1})\cdots Y(v_{k},\beta_{d_{k}}(\gamma)f_{k})\Omega.

for all v1,,vk𝒱QPv_{1},\ldots,v_{k}\in{\mathcal{V}}_{\mathrm{QP}} with conformal dimensions djd_{j} and all fj[z±1]f_{j}\in{\mathbb{C}}[z^{\pm 1}].

Proof.

Uniqueness is clear as the required formula for U0(γ)U^{0}(\gamma) determines its value on 𝒱{\mathcal{V}}. In order to show existence of U0(γ)U^{0}(\gamma), we must show that if a linear combination of vectors of the form Y(v1,f1)Y(vk,fk)ΩY(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega vanishes, then so does the corresponding linear combination of vectors of the form Y(v1,βd1(γ)f1)Y(vk,βdk(γ)fk)ΩY(v^{1},\beta_{d_{1}}(\gamma)f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma)f_{k})\Omega. We use standard ODE techniques.

The exponential map exp:Lie(Mo¨b)Mo¨b\exp:\operatorname{Lie}(\operatorname{M\ddot{o}b})\to\operatorname{M\ddot{o}b} is surjective, so we may choose gddϑLie(Mo¨b)g\tfrac{d}{d\vartheta}\in\operatorname{Lie}(\operatorname{M\ddot{o}b}) such that exp(gddϑ)=γ\exp(g\tfrac{d}{d\vartheta})=\gamma. Let γt=exp(tgddϑ)\gamma_{t}=\exp(tg\tfrac{d}{d\vartheta}) be the corresponding one-parameter subgroup of Mo¨b\operatorname{M\ddot{o}b}. Let v1,,vk𝒱QPv^{1},\ldots,v^{k}\in{\mathcal{V}}_{\mathrm{QP}} and consider the function u:𝒟QPu:{\mathbb{R}}\to{\mathcal{D}}_{\mathrm{QP}} given by

u(t)=Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω.u(t)=Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega.

We will now show that uu extends holomorphically to a neighborhood of {\mathbb{R}} (when 𝒟QP{\mathcal{D}}_{\mathrm{QP}} is given the weak topology induced by 𝒱{\mathcal{V}}^{\prime}), and compute its derivative.

The map Mo¨bPSU(1,1){\mathbb{R}}\to\operatorname{M\ddot{o}b}\cong\mathrm{PSU}(1,1) given by tγtt\mapsto\gamma_{t} extends holomorphically to a neighborhood of {\mathbb{R}} (taking values in complex Möbius transformations of the Riemann sphere PSL2()\cong\mathrm{PSL}_{2}({\mathbb{C}})). For each tt\in{\mathbb{R}}, the Möbius transformation γt\gamma_{t} leaves S1S^{1} invariant, and thus for a sufficiently small neighborhood of {\mathbb{R}} the corresponding Möbius transformations map S1S^{1} into ×{\mathbb{C}}^{\times}. Thus if f[z±1]f\in{\mathbb{C}}[z^{\pm 1}] is a Laurent polynomial, the function ×S1{\mathbb{R}}\times S^{1}\to{\mathbb{C}} given by (t,z)(βd(γt)f)(z)(t,z)\mapsto(\beta_{d}(\gamma_{t})f)(z) extends holomorphically to a neighborhood of ×S1{\mathbb{R}}\times S^{1}. It follows that the map C(S1){\mathbb{R}}\to C^{\infty}(S^{1}) sending tβd(γt)ft\mapsto\beta_{d}(\gamma_{t})f extends holomorphically to a neighborhood of {\mathbb{R}}.

Fix λ𝒱(n)\lambda\in{\mathcal{V}}(n)^{*}. By Lemma 3.4, the expressions λ(Y(v1,f1)Y(vk,fk)Ω)\lambda(Y(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega) are jointly continuous in fjC(S1)f_{j}\in C^{\infty}(S^{1}). Thus for fixed Laurent polynomials f1,,fk[z±1]f_{1},\ldots,f_{k}\in{\mathbb{C}}[z^{\pm 1}], the function

tλ(Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω)t\mapsto\lambda\big{(}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}

extends holomorphically to a neighborhood of {\mathbb{R}}. As this neighborhood is independent of λ\lambda, the function

Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)ΩY(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega

extends holomorphically to a neighborhood of {\mathbb{R}}, as previously claimed.

We now differentiate the above function of tt. A straightforward computation [RTT22, Eqn.​ (3.4)] shows that

ddtβd(γt)f=(d1)dgdϑβd(γt)fgddϑ[βd(γt)f]\frac{d}{dt}\beta_{d}(\gamma_{t})f=(d-1)\tfrac{dg}{d\vartheta}\beta_{d}(\gamma_{t})f-g\tfrac{d}{d\vartheta}[\beta_{d}(\gamma_{t})f] (3.6)

with the derivative taken in C(S1)C^{\infty}(S^{1}). Comparing (3.6) with the commutation relation of Lemma 3.6 we obtain for any λ𝒱\lambda\in{\mathcal{V}}^{\prime}

ddt\displaystyle\frac{d}{dt} λ(Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω)=\displaystyle\lambda\big{(}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}=
=j=1kλ(Y(v1,βd1(γt)f1)Y(vj,ddtβdj(γt)fj)Y(vk,βdk(γt)fk)Ω)\displaystyle=\sum_{j=1}^{k}\lambda\big{(}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{j},\tfrac{d}{dt}\beta_{d_{j}}(\gamma_{t})f_{j})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}
=j=1kλ(Y(v1,βd1(γt)f1)[π(gddϑ),Y(vj,fj)]Y(vk,βdk(γt)fk)Ω)\displaystyle=\sum_{j=1}^{k}\lambda\big{(}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots[\pi(g\tfrac{d}{d\vartheta}),Y(v^{j},f_{j})]\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}
=λ(π(gddϑ)Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω).\displaystyle=\lambda\big{(}\pi(g\tfrac{d}{d\vartheta})Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}.

Since the adjoint operator π(gddϑ)\pi(g\tfrac{d}{d\vartheta})^{*} leaves 𝒱{\mathcal{V}}^{\prime} invariant, we may iterate the above argument to obtain

dmdtm\displaystyle\frac{d^{m}}{dt^{m}} λ(Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω)=\displaystyle\lambda\big{(}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}=
=λ(π(gddϑ)mY(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω).\displaystyle=\lambda\big{(}\pi(g\tfrac{d}{d\vartheta})^{m}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega\big{)}.

Since λ\lambda was arbitrary we have

dmdtm\displaystyle\frac{d^{m}}{dt^{m}} Y(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω=\displaystyle Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega= (3.7)
=π(gddϑ)mY(v1,βd1(γt)f1)Y(vk,βdk(γt)fk)Ω.\displaystyle=\pi(g\tfrac{d}{d\vartheta})^{m}Y(v^{1},\beta_{d_{1}}(\gamma_{t})f_{1})\cdots Y(v^{k},\beta_{d_{k}}(\gamma_{t})f_{k})\Omega.

We now complete the proof of existence of the map U0(γ):𝒱𝒟QPU^{0}(\gamma):{\mathcal{V}}\to{\mathcal{D}}_{\mathrm{QP}}. Suppose that a certain linear combination of vectors of the form Y(v1,f1)Y(vk,fk)ΩY(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega vanishes. That is, suppose we have

i=1Y(vi,1,fi,1)Y(vi,ki,fi,ki)Ω=0\sum_{i=1}^{\ell}Y(v^{i,1},f_{i,1})\cdots Y(v^{i,k_{i}},f_{i,k_{i}})\Omega=0

for ,ki0\ell,k_{i}\in{\mathbb{Z}}_{\geq 0}, vi,jv^{i,j} quasiprimary vectors with conformal dimension di,jd_{i,j}, and fi,jf_{i,j} Laurent polynomials. Then, by the above, the function

ti=1Y(vi,1,βdi,1(γt)fi,1)Y(vi,ki,βdi,ki(γt)fi,ki)Ωt\mapsto\sum_{i=1}^{\ell}Y(v^{i,1},\beta_{d_{i,1}}(\gamma_{t})f_{i,1})\cdots Y(v^{i,k_{i}},\beta_{d_{i,k_{i}}}(\gamma_{t})f_{i,k_{i}})\Omega (3.8)

extends holomorphically to a neighborhood of {\mathbb{R}}, and by (3.7) the derivatives of (3.8) at t=0t=0 are given by

dmdtm\displaystyle\frac{d^{m}}{dt^{m}} i=1Y(vi,1,βdi,1(γt)fi,1)Y(vi,ki,βdi,ki(γt)fi,ki)Ω|t=0=\displaystyle\left.\sum_{i=1}^{\ell}Y(v^{i,1},\beta_{d_{i,1}}(\gamma_{t})f_{i,1})\cdots Y(v^{i,k_{i}},\beta_{d_{i,k_{i}}}(\gamma_{t})f_{i,k_{i}})\Omega\,\right|_{t=0}=
=π(gddϑ)mi=1Y(vi,1,fi,1)Y(vi,ki,fi,ki)Ω\displaystyle=\pi(g\tfrac{d}{d\vartheta})^{m}\sum_{i=1}^{\ell}Y(v^{i,1},f_{i,1})\cdots Y(v^{i,k_{i}},f_{i,k_{i}})\Omega
=0.\displaystyle=0.

Since the Taylor series of (3.8) at t=0t=0 is identically zero, the function vanishes identically. In particular, specializing to t=1t=1 yields

i=1Y(vi,1,βdi,1(γ)fi,1)Y(vi,ki,βdi,ki(γ)fi,ki)Ω=0.\sum_{i=1}^{\ell}Y(v^{i,1},\beta_{d_{i,1}}(\gamma)f_{i,1})\cdots Y(v^{i,k_{i}},\beta_{d_{i,k_{i}}}(\gamma)f_{i,k_{i}})\Omega=0.

We conclude that the desired map U0(γ)U^{0}(\gamma) is well-defined, as required. ∎

We now address the problem of extending U0(γ)U^{0}(\gamma) to an endomorphism of 𝒟QP{\mathcal{D}}_{\mathrm{QP}}.

Lemma 3.8.

Let 𝒱{\mathcal{V}} be a Möbius vertex algebra which is generated by its quasiprimary vectors. Then for any γMo¨b\gamma\in\operatorname{M\ddot{o}b} there exists a unique linear map U(γ)(𝒟QP)U(\gamma)\in{\mathcal{L}}({\mathcal{D}}_{\mathrm{QP}}) such that

U(γ)Y(v1,f1)Y(vk,fk)Ω=Y(v1,βd1(γ)f1)Y(vk,βdk(γ)fk)Ω.U(\gamma)Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega=Y(v_{1},\beta_{d_{1}}(\gamma)f_{1})\cdots Y(v_{k},\beta_{d_{k}}(\gamma)f_{k})\Omega.

for all v1,,vk𝒱QPv_{1},\ldots,v_{k}\in{\mathcal{V}}_{\mathrm{QP}} with conformal dimensions djd_{j} and all fjC(S1)f_{j}\in C^{\infty}(S^{1}).

Proof.

Let Φ𝒟QP\Phi\in{\mathcal{D}}_{\mathrm{QP}}, and recall that 𝒟QP𝒱^=n=0𝒱(n){\mathcal{D}}_{\mathrm{QP}}\subset\widehat{\mathcal{V}}=\prod_{n=0}^{\infty}{\mathcal{V}}(n). Thus we may canonically write Φ=n=0un\Phi=\sum_{n=0}^{\infty}u^{n} with un𝒱(n)u^{n}\in{\mathcal{V}}(n) (and the sum converging in the weak topology induced by 𝒱{\mathcal{V}}^{\prime}). We would like to define U(γ)Φ=n=0U0(γ)unU(\gamma)\Phi=\sum_{n=0}^{\infty}U^{0}(\gamma)u^{n}, but first must check convergence of the sum. It suffices to consider a vector Φ=Y(v1,f1)Y(vk,fk)Ω\Phi=Y(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega with vj𝒱QPv_{j}\in{\mathcal{V}}_{\mathrm{QP}} and fjC(S1)f_{j}\in C^{\infty}(S^{1}). Recall from Lemma 3.2 that the continuous map multilinear map C(S1)k𝒟QPC^{\infty}(S^{1})^{k}\to{\mathcal{D}}_{\mathrm{QP}} given by

(f1,,fk)Y(v1,f1)Y(vk,fk)Ω(f_{1},\ldots,f_{k})\mapsto Y(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega

takes values in n=0𝒱(n;v1,,vk)\prod_{n=0}^{\infty}{\mathcal{V}}(n;v^{1},\ldots,v^{k}) with 𝒱(n;v1,,vk){\mathcal{V}}(n;v^{1},\ldots,v^{k}) a finite-dimensional subspace of 𝒱(n){\mathcal{V}}(n). Thus by the universal property of the projective tensor product π\otimes_{\pi} [Trè67, Prop. 43.4] we have a continuous linear map

C(S1)ππC(S1)n=0𝒱(n;v1,,vk).C^{\infty}(S^{1})\otimes_{\pi}\cdots\otimes_{\pi}C^{\infty}(S^{1})\to\prod_{n=0}^{\infty}{\mathcal{V}}(n;v^{1},\ldots,v^{k}).

As n=0𝒱(n;v1,,vk)\prod_{n=0}^{\infty}{\mathcal{V}}(n;v^{1},\ldots,v^{k}) is complete, by [Trè67, Thm. 5.2] (see also Appendix B for the completion of topological vector spaces) we may extend this map to a continuous linear map

C(S1)^π^πC(S1)n=0𝒱(n;v1,,vk)C^{\infty}(S^{1})\hat{\otimes}_{\pi}\cdots\hat{\otimes}_{\pi}C^{\infty}(S^{1})\to\prod_{n=0}^{\infty}{\mathcal{V}}(n;v^{1},\ldots,v^{k})

where ^π\hat{\otimes}_{\pi} is the completed projective tensor product (see [Trè67, §43]). We have a natural isomorphism of topological vector spaces

C(S1)^π^πC(S1)C((S1)k)C^{\infty}(S^{1})\hat{\otimes}_{\pi}\cdots\hat{\otimes}_{\pi}C^{\infty}(S^{1})\cong C^{\infty}((S^{1})^{k})

by [Trè67, Thm. 56.1] (extended to the manifold S1S^{1} via partition of unity). Thus we conclude that there exists a continuous linear map

S:C((S1)k)n=0𝒱(n;v1,,vk)S:C^{\infty}((S^{1})^{k})\to\prod_{n=0}^{\infty}{\mathcal{V}}(n;v^{1},\ldots,v^{k})

characterized by

S(f1ππfk)=Y(v1,f1)Y(vk,fk)Ω.S(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k})=Y(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega.

Now fix v1,,vk𝒱v^{1},\ldots,v^{k}\in{\mathcal{V}} and f1,,fkC(S1)f_{1},\ldots,f_{k}\in C^{\infty}(S^{1}), and consider

Y(v1,f1)Y(vk,fk)Ω=Φ=n=0un,Y(v^{1},f_{1})\cdots Y(v^{k},f_{k})\Omega=\Phi=\sum_{n=0}^{\infty}u^{n},

again with un𝒱(n)u^{n}\in{\mathcal{V}}(n). Let V(γ):C((S1)k)C((S1)k)V(\gamma):C^{\infty}((S^{1})^{k})\to C^{\infty}((S^{1})^{k}) be the continuous linear map such that

V(γ)g1(z1)gk(zk)=(βd1(γ)g1)(z1)(βd1(γ)gk)(zk)V(\gamma)g_{1}(z_{1})\cdots g_{k}(z_{k})=(\beta_{d_{1}}(\gamma)g_{1})(z_{1})\cdots(\beta_{d_{1}}(\gamma)g_{k})(z_{k})

for all gjC(S1)g_{j}\in C^{\infty}(S^{1}). For nn\in{\mathbb{Z}}, let Pn:C((S1)k)C((S1)k)P_{n}:C^{\infty}((S^{1})^{k})\to C^{\infty}((S^{1})^{k}) be the natural projection onto the closed span of monomials z1n1zknkz_{1}^{n_{1}}\cdots z_{k}^{n_{k}} with n1++nk=nn_{1}+\cdots+n_{k}=-n (whose kernel is spanned by monomials with n1++nknn_{1}+\cdots+n_{k}\neq-n). For F(z1,,zk)=f1(z1)fk(zk)F(z_{1},\ldots,z_{k})=f_{1}(z_{1})\cdots f_{k}(z_{k}), by construction we have S(PnF)=unS(P_{n}F)=u^{n} and S(V(γ)PnF)=U0(γ)unS(V(\gamma)P_{n}F)=U^{0}(\gamma)u^{n}, where U0U^{0} is defined in Lemma 3.7. Since nPn=id\sum_{n\in{\mathbb{Z}}}P_{n}=\operatorname{id} (with convergence pointwise as operators on C((S1)k)C^{\infty}((S^{1})^{k})) and both SS and V(γ)V(\gamma) are continuous, we have convergence of the sum

n=0U0(γ)un=n=0S(V(γ)PnF)\sum_{n=0}^{\infty}U^{0}(\gamma)u^{n}=\sum_{n=0}^{\infty}S(V(\gamma)P_{n}F)

to S(V(γ)F)=Y(v1,βd1(γ)f1)Y(vk,βdk(γ)fk)ΩS(V(\gamma)F)=Y(v_{1},\beta_{d_{1}}(\gamma)f_{1})\cdots Y(v_{k},\beta_{d_{k}}(\gamma)f_{k})\Omega.

As the action of U0(γ)U^{0}(\gamma) on un𝒱(n)u^{n}\in{\mathcal{V}}(n) is well-defined by Lemma 3.7 and does not depend on the choice of {vj}\{v^{j}\}, we have obtained both a well-defined map U(γ)U(\gamma) given by U(γ)Φ=n=0U0(γ)unU(\gamma)\Phi=\sum_{n=0}^{\infty}U^{0}(\gamma)u^{n} along with the required covariance relation. ∎

Since βd(γ1)βd(γ2)=βd(γ1γ2)\beta_{d}(\gamma_{1})\beta_{d}(\gamma_{2})=\beta_{d}(\gamma_{1}\circ\gamma_{2}), the maps U(γ)U(\gamma) furnish a representation of Mo¨b\operatorname{M\ddot{o}b} on 𝒟QP{\mathcal{D}}_{\mathrm{QP}}. For v𝒱QPv\in{\mathcal{V}}_{\mathrm{QP}} the operator-valued distribution Y(v,f)|𝒟QPY(v,f)|_{{\mathcal{D}}_{\mathrm{QP}}} is evidently covariant with respect to this representation.

Theorem 3.9.

Let 𝒱{\mathcal{V}} be a Möbius vertex algebra, let S𝒱S\subset{\mathcal{V}} be a set of quasiprimary vectors that generate 𝒱{\mathcal{V}} as a vertex algebra, and let

𝒟S=span{Y(v1,f1)Y(vk,fk)Ω|k0,vjS,fjC(S1)}.{\mathcal{D}}_{S}=\operatorname{span}\{Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Omega\,|\,k\in{\mathbb{Z}}_{\geq 0},v_{j}\in S,f_{j}\in C^{\infty}(S^{1})\}.

Let S={Y(v,f)|𝒟S|vS,fC(S1)}{\mathcal{F}}_{S}=\{Y(v,f)|_{{\mathcal{D}}_{S}}\,|\,v\in S,f\in C^{\infty}(S^{1})\} and for γMo¨b\gamma\in\operatorname{M\ddot{o}b} let US(γ)=U(γ)|SU_{S}(\gamma)=U(\gamma)|_{S}. Then (S,𝒟S,US,Ω)({\mathcal{F}}_{S},{\mathcal{D}}_{S},U_{S},\Omega) is a (not-necessarily-unitary) Möbius-covariant Wightman CFT.

Proof.

We have a family of operator-valued distributions S{\mathcal{F}}_{S} on 𝒟S𝒟QP𝒱^{\mathcal{D}}_{S}\subset{\mathcal{D}}_{\mathrm{QP}}\subset\widehat{\mathcal{V}}. Note that 𝒱𝒟S{\mathcal{V}}\subset{\mathcal{D}}_{S} since SS generates 𝒱{\mathcal{V}}. By Lemma 3.4 we have 𝒱𝒟,S{\mathcal{V}}^{\prime}\subset{\mathcal{D}}_{{\mathcal{F}},S}^{*}, where we note that it suffices to check continuity of (Y(v1,f1)Y(vk,fk)Φ,u)\left(Y(v_{1},f_{1})\cdots Y(v_{k},f_{k})\Phi,u^{\prime}\right) in the special case Φ=Ω\Phi=\Omega since 𝒟S{\mathcal{D}}_{S} is generated from Ω\Omega by S{\mathcal{F}}_{S}. Hence 𝒟,S{\mathcal{D}}_{{\mathcal{F}},S}^{*} separates points, as 𝒱{\mathcal{V}}^{\prime} separates points in 𝒱^\widehat{\mathcal{V}}, and so S{\mathcal{F}}_{S} acts regularly. The subspace 𝒟𝒮{\mathcal{D}}_{\mathcal{S}} is invariant under UU by Lemma 3.8, and by the same lemma the fields Y(v,f)Y(v,f) are Möbius covariant, which verifies the first axiom of a Wightman CFT.

We now check the locality axiom. Let v1,v2Sv_{1},v_{2}\in S, let u𝒱u\in{\mathcal{V}} and let u𝒱u^{\prime}\in{\mathcal{V}}^{\prime}. By the vertex algebra locality axiom, the formal distribution (z1z2)N([Y(v1,z1),Y(v2,z2)]u,u)(z_{1}-z_{2})^{N}\left([Y(v_{1},z_{1}),Y(v_{2},z_{2})]u,u^{\prime}\right) vanishes for NN sufficiently large, and thus the corresponding distribution (f1,f2)([Y(v1,f1),Y(v2,f2)]u,u)(f_{1},f_{2})\mapsto\left([Y(v_{1},f_{1}),Y(v_{2},f_{2})]u,u^{\prime}\right) is supported on the diagonal z1=z2z_{1}=z_{2} (see [Kac98, Cor. 2.2] and [CKLW18, Prop. A.1]). Hence when f1f_{1} and f2f_{2} have disjoint support we have

[Y(v1,f1),Y(v2,f2)]u=0forallu𝒱.[Y(v_{1},f_{1}),Y(v_{2},f_{2})]u=0\quad\mathrm{for\ all}\ u\in{\mathcal{V}}.

That is,

[Y(v1,f1),Y(v2,f2)]Y(a1,g1)Y(ak,gk)Ω=0[Y(v_{1},f_{1}),Y(v_{2},f_{2})]Y(a_{1},g_{1})\cdots Y(a_{k},g_{k})\Omega=0

for all ajSa_{j}\in S and gj[z±1]g_{j}\in{\mathbb{C}}[z^{\pm 1}]. By the joint continuity of such expressions in gjg_{j} (which shows that 𝒱{\mathcal{V}} is {\mathcal{F}}-weakly dense in 𝒟S{\mathcal{D}}_{S}) and the cyclicity of Ω\Omega, we see that [Y(v1,f1),Y(v2,f2)]Φ=0[Y(v_{1},f_{1}),Y(v_{2},f_{2})]\Phi=0 for all Φ𝒟S\Phi\in{\mathcal{D}}_{S}, and thus the locality axiom holds.

The vacuum axiom holds by construction, and the spectrum condition holds by Lemma 2.11. ∎

3.2 From Wightman CFTs to vertex algebras

Let {\mathcal{F}} be a Wightman CFT with domain 𝒟{\mathcal{D}}, with vacuum vector Ω\Omega and representation U:Mo¨b(𝒟)U:\operatorname{M\ddot{o}b}\to{\mathcal{L}}({\mathcal{D}}). Let 𝒱(n)𝒟{\mathcal{V}}(n)\subset{\mathcal{D}} be the finite energy subspace

𝒱(n)=span{φ1(ej1)φk(ejk)Ω|k0,ji=n,φi},{\mathcal{V}}(n)=\operatorname{span}\{\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega\,|\,k\in{\mathbb{Z}}_{\geq 0},\sum j_{i}=-n,\varphi_{i}\in{\mathcal{F}}\},

where ej(z)=zje_{j}(z)=z^{j}, and let 𝒱=n0𝒱(n)𝒟{\mathcal{V}}=\bigoplus_{n\geq 0}{\mathcal{V}}(n)\subset{\mathcal{D}}. Note that when n<0n<0 we have 𝒱(n)=0{\mathcal{V}}(n)=0 by the spectrum condition of a Wightman CFT, and 𝒱{\mathcal{V}} is {\mathcal{F}}-strongly dense in 𝒟{\mathcal{D}}.

We will show that 𝒱{\mathcal{V}} carries the structure of a Möbius vertex algebra generated by the point-like quasiprimary fields corresponding to {\mathcal{F}}. For φ\varphi\in{\mathcal{F}} with conformal dimension dd, the corresponding point-like field is a formal sum

φ^(z)=nφ(en)znd.\hat{\varphi}(z)=\sum_{n\in{\mathbb{Z}}}\varphi(e_{n})z^{-n-d}.

The key steps are to establish the vertex algebra locality condition

(zw)N[φ^(z),ψ^(w)]=0(z-w)^{N}[\hat{\varphi}(z),\hat{\psi}(w)]=0

for NN sufficiently large, as well as differentiating the representation UU to a representation of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}} for which we have the infinitesimal Möbius covariance condition

[Lm,φ^(z)]=(zm+1ddz+(m+1)zmd)φ^(z),m=1,0,1.[L_{m},\hat{\varphi}(z)]=\big{(}z^{m+1}\tfrac{d}{dz}+(m+1)z^{m}d\big{)}\hat{\varphi}(z),\quad m=-1,0,1.

From there, we will invoke general results that say that families of covariant local fields produce vertex algebras (see [Kac98, Thm. 4.5] for the case of vertex algebras, or more specifically [RTT22, Thm. A.1] for a slight variant for Möbius vertex algebras).

We begin by establishing Möbius covariance.

Lemma 3.10.

There is a unique representation π:Lie(Mo¨b)(𝒟)\pi:\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}}\to{\mathcal{L}}({\mathcal{D}}) such that for all φ\varphi\in{\mathcal{F}} with conformal dimension dd and all gddϑLie(Mo¨b)g\tfrac{d}{d\vartheta}\in\operatorname{Lie}(\operatorname{M\ddot{o}b}) we have π(gddϑ)Ω=0\pi(g\tfrac{d}{d\vartheta})\Omega=0 and

[π(gddϑ),φ(f)]=φ((d1)dgdϑgdfdϑ).[\pi(g\tfrac{d}{d\vartheta}),\varphi(f)]=\varphi\big{(}(d-1)\tfrac{dg}{d\vartheta}-g\tfrac{df}{d\vartheta}\big{)}.
Proof.

Uniqueness of such a representation follows immediately from the cyclicity of the vacuum (W4). Let gddϑLie(Mo¨b)g\tfrac{d}{d\vartheta}\in\operatorname{Lie}(\operatorname{M\ddot{o}b}), and let γtMo¨b\gamma_{t}\in\operatorname{M\ddot{o}b} be the associated one-parameter group. We have

U(γt)φ1(f1)φn(fn)Ω=φ1(βd1(γt)f1)φn(βdn(γt)fn)Ω,U(\gamma_{t})\varphi_{1}(f_{1})\cdots\varphi_{n}(f_{n})\Omega=\varphi_{1}(\beta_{d_{1}}(\gamma_{t})f_{1})\cdots\varphi_{n}(\beta_{d_{n}}(\gamma_{t})f_{n})\Omega,

where did_{i} is the conformal dimension of φi\varphi_{i}\in{\mathcal{F}}. The derivative of βd(γt)\beta_{d}(\gamma_{t}) is given (as in [RTT22, Eqn.​ (3.4)]) by

ddt|t=0βd(γt)f=(d1)dgdϑfgdfdϑ,\left.\frac{d}{dt}\right|_{t=0}\beta_{d}(\gamma_{t})f=(d-1)\tfrac{dg}{d\vartheta}f-g\tfrac{df}{d\vartheta},

with the derivative taken in the space of smooth functions on S1S^{1}.

Give 𝒟{\mathcal{D}} the {\mathcal{F}}-strong topology. Since expressions φ1(f1)φn(fn)Ω\varphi_{1}(f_{1})\cdots\varphi_{n}(f_{n})\Omega are jointly continuous in the fjf_{j}, we have

ddt|t=0U(γt)φ1(f1)φn(fn)Ω=j=1nφ1(f1)φj((dj1)dgdϑfjgdfjdϑ)φn(fn)Ω.\left.\frac{d}{dt}\right|_{t=0}U(\gamma_{t})\varphi_{1}(f_{1})\cdots\varphi_{n}(f_{n})\Omega=\sum_{j=1}^{n}\varphi_{1}(f_{1})\cdots\varphi_{j}\big{(}(d_{j}-1)\tfrac{dg}{d\vartheta}f_{j}-g\tfrac{df_{j}}{d\vartheta}\big{)}\cdots\varphi_{n}(f_{n})\Omega. (3.9)

In particular, for every Φ𝒟\Phi\in{\mathcal{D}} the expression U(γt)ΦU(\gamma_{t})\Phi is differentiable at t=0t=0, and we define π(gddϑ)Φ=ddt|t=0U(γt)Φ\pi(g\tfrac{d}{d\vartheta})\Phi=\left.\tfrac{d}{dt}\right|_{t=0}U(\gamma_{t})\Phi. We have π(gddϑ)Ω=0\pi(g\tfrac{d}{d\vartheta})\Omega=0 by the Möbius invariance of the vacuum, and from (3.9) we obtain the desired commutation relation for [π(gddϑ),φ(f)][\pi(g\tfrac{d}{d\vartheta}),\varphi(f)]. A direct calculation shows that π\pi is a Lie algebra representation. ∎

Recalling that Lm=π(ieimϑddϑ)L_{m}=\pi(-{\mathrm{i}}{\mathrm{e}}^{{\mathrm{i}}m\vartheta}\tfrac{d}{d\vartheta}) for m=1,0,1m=-1,0,1, one can apply Lemma 3.10 term-by-term to the modes of φ^(z)\hat{\varphi}(z) to deduce the infinitesimal covariance relation

[Lm,φ^(z)]=(zm+1ddz+(m+1)zmd)φ^(z).[L_{m},\hat{\varphi}(z)]=\big{(}z^{m+1}\tfrac{d}{dz}+(m+1)z^{m}d\big{)}\hat{\varphi}(z). (3.10)

We now turn our attention to establishing the vertex algebra locality condition. Recall that 𝒱{\mathcal{V}}^{\prime} denotes n=0𝒱(n)\bigoplus_{n=0}^{\infty}{\mathcal{V}}(n)^{*}; that is, the space of linear functionals on 𝒱{\mathcal{V}} that are supported on finitely many 𝒱(n){\mathcal{V}}(n). By abuse of notation we write 𝒟𝒱{\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}^{\prime} for the subspace of 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} consisting of linear functionals λ\lambda such that λ|𝒱𝒱\lambda|_{\mathcal{V}}\in{\mathcal{V}}^{\prime}, and similarly for 𝒟𝒱(n){\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n)^{*}. By Lemma 2.11 𝒟𝒱(n){\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n)^{*} separates points in 𝒱(n){\mathcal{V}}(n), and so 𝒟𝒱{\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}^{\prime} separates points in 𝒱{\mathcal{V}}. The endomorphism L1=π(ieiϑddϑ)L_{-1}=\pi(-{\mathrm{i}}{\mathrm{e}}^{-{\mathrm{i}}\vartheta}\tfrac{d}{d\vartheta}) of (3.10) and Lemma 3.10 gives an endomorphism of 𝒟{\mathcal{D}} by the lemma. Moreover, the adjoint (transpose) operator L1L_{-1}^{*} leaves 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} invariant, becauase if λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}, then we have by the same lemma

(L1λ)\displaystyle(L_{-1}^{*}\lambda) (φ1(f1)φk(fk)Ω)=\displaystyle(\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega)=
=λ(j=1kφ1(f1)φj((dj1)eiϑfj+ieiϑdfjdϑ)φk(fk)Ω),\displaystyle=\lambda\left(\sum_{j=1}^{k}\varphi_{1}(f_{1})\cdots\varphi_{j}\big{(}-(d_{j}-1){\mathrm{e}}^{-{\mathrm{i}}\vartheta}f_{j}+{\mathrm{i}}{\mathrm{e}}^{-{\mathrm{i}}\vartheta}\tfrac{df_{j}}{d\vartheta}\big{)}\cdots\varphi_{k}(f_{k})\Omega\right),

which depends continuously on the fjf_{j}, so L1λ𝒟L_{-1}^{*}\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}. Hence L1L_{-1}^{*} leaves 𝒟𝒱{\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}^{\prime} invariant, mapping 𝒟𝒱(n){\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n)^{*} into 𝒟𝒱(n1){\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}(n-1)^{*}.

If λ\lambda is a linear functional on a vector space VV and A(z1,,zk)A(z_{1},\ldots,z_{k}) is a formal series with coefficients in VV, then we write λ(A(z1,,zk))\lambda(A(z_{1},\ldots,z_{k})) for the corresponding formal series with coefficients in {\mathbb{C}}.

Lemma 3.11.

Let φ1,φ2\varphi_{1},\varphi_{2}\in{\mathcal{F}} with conformal dimensions d1d_{1} and d2d_{2}, respectively. Then for every λ𝒱\lambda\in{\mathcal{V}}^{\prime} the formal series

(z1z2)d1+d2λ(φ^1(z1)φ^2(z2)Ω)(z_{1}-z_{2})^{d_{1}+d_{2}}\lambda\big{(}\hat{\varphi}_{1}(z_{1})\hat{\varphi}_{2}(z_{2})\Omega\big{)}

is a polynomial in z1z_{1} and z2z_{2} after expanding (z1z2)d1+d2(z_{1}-z_{2})^{d_{1}+d_{2}} using the binomial theorem.

Proof.

We use standard vertex algebra arguments which go through in the present context. From the positivity of the energy and the L0L_{0}- and L1L_{-1}-commutation relations (3.10), we can deduce (as in the proof of [RTT22, Thm. 3.11]) that φ^2(z2)Ω\hat{\varphi}_{2}(z_{2})\Omega has only non-negative powers of z2z_{2}, and if u:=φ^2(z2)Ωz2=0u:=\hat{\varphi}_{2}(z_{2})\Omega\mid_{z_{2}=0} is the constant term, then u𝒱(d2)u\in{\mathcal{V}}(d_{2}). The formal power series ez2L1u{\mathrm{e}}^{z_{2}L_{-1}}u and φ^(z2)Ω\hat{\varphi}(z_{2})\Omega both solve the initial value problem ddz2F(z2)=L1F(z2)\tfrac{d}{dz_{2}}F(z_{2})=L_{-1}F(z_{2}) with F(0)=uF(0)=u. This initial value problem has a unique solution in 𝒱[[z2]]{\mathcal{V}}[[z_{2}]], and we conclude φ^2(z2)Ω=ez2L1u\hat{\varphi}_{2}(z_{2})\Omega={\mathrm{e}}^{z_{2}L_{-1}}u as formal series.

Similarly, we consider the formal series in z1±1z_{1}^{\pm 1} and z2z_{2} given by ez2L1φ^1(z1)ez2L1{\mathrm{e}}^{-z_{2}L_{-1}}\hat{\varphi}_{1}(z_{1}){\mathrm{e}}^{z_{2}L_{-1}}. It satisfies the initial value problem ddz2F(z1,z2)=[L1,F(z1,z2)]\tfrac{d}{dz_{2}}F(z_{1},z_{2})=-[L_{-1},F(z_{1},z_{2})] with F(z1,0)=φ^1(z1)F(z_{1},0)=\hat{\varphi}_{1}(z_{1}). Taking each coefficient of z1mz_{1}^{m} separately, it is straightforward to see that this initial value problem has a unique solution in End(𝒱)[[z1±1,z2]]\operatorname{End}({\mathcal{V}})[[z_{1}^{\pm 1},z_{2}]]. Let ιz1,z2φ^1(z1z2)\iota_{z_{1},z_{2}}\hat{\varphi}_{1}(z_{1}-z_{2}) denote the series in End(𝒱)[[z1±1,z2]]\operatorname{End}({\mathcal{V}})[[z_{1}^{\pm 1},z_{2}]] obtained by expanding each term (z1z2)m(z_{1}-z_{2})^{m} as a binomial series with positive powers of z2z_{2}. This series satisfies the same initial value problem, and so we have

ez2L1φ^1(z1)ez2L1=ιz1,z2φ^1(z1z2).{\mathrm{e}}^{-z_{2}L_{-1}}\hat{\varphi}_{1}(z_{1}){\mathrm{e}}^{z_{2}L_{-1}}=\iota_{z_{1},z_{2}}\hat{\varphi}_{1}(z_{1}-z_{2}).

Putting the two calculations together, we obtain an identity of formal series

φ^1(z1)φ^2(z2)Ω=ez2L1ιz1,z2φ^1(z1z2)u.\hat{\varphi}_{1}(z_{1})\hat{\varphi}_{2}(z_{2})\Omega={\mathrm{e}}^{z_{2}L_{-1}}\iota_{z_{1},z_{2}}\hat{\varphi}_{1}(z_{1}-z_{2})u.

Hence

λ(φ^1(z1)φ^2(z2)Ω)=(ez2L1λ)(ιz1,z2φ^1(z1z2)u).\lambda\big{(}\hat{\varphi}_{1}(z_{1})\hat{\varphi}_{2}(z_{2})\Omega\big{)}=({\mathrm{e}}^{z_{2}L_{-1}^{*}}\lambda)\big{(}\iota_{z_{1},z_{2}}\hat{\varphi}_{1}(z_{1}-z_{2})u\big{)}.

As L1L_{-1}^{*} maps 𝒱(n){\mathcal{V}}(n)^{*} into 𝒱(n1){\mathcal{V}}(n-1)^{*}, it acts nilpotently on λ\lambda and the sum defining ez1L1λ{\mathrm{e}}^{z_{1}L_{-1}^{*}}\lambda is finite.

Consider a term of this sum, which is of the form ((L1)mλ)(φ^1(z1z2)u)((L_{-1}^{*})^{m}\lambda)(\hat{\varphi}_{1}(z_{1}-z_{2})u). It suffices to prove the lemma for λ𝒱(d)\lambda\in{\mathcal{V}}(d)^{*} and then take linear combinations, in which case there is at most one non-zero term in the sum defining this expression. That is, if we write φ^1(z)=φ^1,nznd1\hat{\varphi}_{1}(z)=\sum\hat{\varphi}_{1,n}z^{-n-d_{1}} then

((L1)mλ)(φ^1(z1z2)u)=(z1z2)d1d2+dmλ(φ^1,d2d+mu).((L_{-1}^{*})^{m}\lambda)(\hat{\varphi}_{1}(z_{1}-z_{2})u)=(z_{1}-z_{2})^{-d_{1}-d_{2}+d-m}\lambda(\hat{\varphi}_{1,d_{2}-d+m}u).

Since this term is non-zero only when mdm\leq d, we have that

(z1z2)d1+d2ιz1,z2((L1)mλ)(φ^1(z1z2)u)=ιz1,z2C(z1z2)dm(z_{1}-z_{2})^{d_{1}+d_{2}}\iota_{z_{1},z_{2}}((L_{-1}^{*})^{m}\lambda)(\hat{\varphi}_{1}(z_{1}-z_{2})u)=\iota_{z_{1},z_{2}}C(z_{1}-z_{2})^{d-m}

for a constant CC, which is a polynomial in z1z_{1} and z2z_{2}. We conclude that

(z1z2)d1+d2ιz1,z2(ez2L1λ)(φ^1(z1z2)u)(z_{1}-z_{2})^{d_{1}+d_{2}}\iota_{z_{1},z_{2}}({\mathrm{e}}^{z_{2}L_{-1}^{*}}\lambda)(\hat{\varphi}_{1}(z_{1}-z_{2})u)

is a polynomial in z1z_{1} and z2z_{2}, and we are done. ∎

Lemma 3.12.

Let {\mathcal{F}} be a Möbius-covariant Wightman CFT, and let φ1,φ2\varphi_{1},\varphi_{2}\in{\mathcal{F}}. Then φ^1\hat{\varphi}_{1} and φ^2\hat{\varphi}_{2} are local in the sense of vertex algebras.

Proof.

Let X:C(S1)×C(S1)End(𝒟)X:C^{\infty}(S^{1})\times C^{\infty}(S^{1})\to\operatorname{End}({\mathcal{D}}) be the operator-valued distribution corresponding to the formal series (z1z2)d1+d2[φ^1(z1),φ^2(z2)](z_{1}-z_{2})^{d_{1}+d_{2}}[\hat{\varphi}_{1}(z_{1}),\hat{\varphi}_{2}(z_{2})] after expanding out the binomial (z1z2)d1+d2(z_{1}-z_{2})^{d_{1}+d_{2}}. More precisely, we first define XX on pairs of functions (en,em)(e_{n},e_{m}), where en(z)=zne_{n}(z)=z^{n}, by

(z1z2)d1+d2[φ^1(z1),φ^2(z2)]=n,mX(en,em)z1n1z2m1,(z_{1}-z_{2})^{d_{1}+d_{2}}[\hat{\varphi}_{1}(z_{1}),\hat{\varphi}_{2}(z_{2})]=\sum_{n,m\in{\mathbb{Z}}}X(e_{n},e_{m})z_{1}^{-n-1}z_{2}^{-m-1},

and these coefficients lie in End(𝒱)\operatorname{End}({\mathcal{V}}). However expanding (z1z2)d1+d2(z_{1}-z_{2})^{d_{1}+d_{2}} we see that XX is a (finite) linear combination of distributions of the form

(f,g)[φ1(enf),φ2(emg)],(f,g)\mapsto[\varphi_{1}(e_{n}\cdot f),\varphi_{2}(e_{m}\cdot g)], (3.11)

which extends to a genuine distribution X:C(S1)×C(S1)End(𝒟)X:C^{\infty}(S^{1})\times C^{\infty}(S^{1})\to\operatorname{End}({\mathcal{D}}) as claimed. Moreover, we see from this formula that X(f,g)=0X(f,g)=0 when ff and gg have disjoint support, i.e.​ the support of XX is contained in the diagonal of S1×S1S^{1}\times S^{1}.

Let λ𝒟𝒱\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}^{\prime}, and note that since λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*} the distribution

(f,g)λ(X(f,g)Ω)(f,g)\mapsto\lambda(X(f,g)\Omega)

is indeed continuous in ff and gg. Applying Lemma 3.11 twice, we see that this distribution, which corresponds to the formal series (z1z2)d1+d2λ([φ^1(z1),φ^2(z2)]Ω)(z_{1}-z_{2})^{d_{1}+d_{2}}\lambda([\hat{\varphi}_{1}(z_{1}),\hat{\varphi}_{2}(z_{2})]\Omega), is given by integration against a trigonometric polynomial. As noted above this distribution (and hence the corresponding polynomial) has support contained in the diagonal of S1×S1S^{1}\times S^{1}, and thus must be identically zero. Since 𝒟𝒱{\mathcal{D}}_{\mathcal{F}}^{*}\cap{\mathcal{V}}^{\prime} separates points in 𝒱{\mathcal{V}} by Lemma 2.11 we conclude that X(en,em)Ω=0X(e_{n},e_{m})\Omega=0 for all n,mn,m\in{\mathbb{Z}}. As X(f,g)ΩX(f,g)\Omega is {\mathcal{F}}-weakly continuous in ff and gg, this implies that X(f,g)Ω=0X(f,g)\Omega=0 for all f,gC(S1)f,g\in C^{\infty}(S^{1}).

Recall that XX is a linear combination of distributions of the form (3.11). Hence if ff and gg are supported in an open, non-dense interval II of the circle, then the Reeh-Schlieder property (Corollary A.3) implies that X(f,g)=0X(f,g)=0. Now choose three intervals that cover S1S^{1} such that the union of any two is contained inside some interval, and let {χi}\{\chi_{i}\} be a partition of unity subordinate to this cover. Then X(f,g)=i,j=13X(fχi,gχj)=0X(f,g)=\sum_{i,j=1}^{3}X(f\chi_{i},g\chi_{j})=0 for arbitrary f,gC(S1)f,g\in C^{\infty}(S^{1}). In particular X(en,em)=0X(e_{n},e_{m})=0 for all n,mn,m\in{\mathbb{Z}}, and we conclude that the formal series (z1z2)d1+d2[φ^1(z1),φ^2(z2)](z_{1}-z_{2})^{d_{1}+d_{2}}[\hat{\varphi}_{1}(z_{1}),\hat{\varphi}_{2}(z_{2})] is identically zero, as desired. ∎

We can now state and prove one of our main results, constructing a Möbius vertex algebra from a Wightman theory.

Theorem 3.13.

Let {\mathcal{F}} be a (not-necessarily-unitary) Möbius-covariant Wightman CFT on S1S^{1} with domain 𝒟{\mathcal{D}}, and let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} be given by

𝒱=span{φ1(ej1)φk(ejk)Ω|k0,ji,φi}.{\mathcal{V}}=\operatorname{span}\{\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega\,|\,k\in{\mathbb{Z}}_{\geq 0},\,j_{i}\in{\mathbb{Z}},\,\varphi_{i}\in{\mathcal{F}}\}.

Then there is a unique structure of Möbius vertex algebra on 𝒱{\mathcal{V}} such that for every φ\varphi\in{\mathcal{F}} with conformal dimension dd there is a quasiprimary vφ𝒱(d)v_{\varphi}\in{\mathcal{V}}(d) such that φ^(z)=Y(vφ,z)End(𝒱)[[z±1]]\hat{\varphi}(z)=Y(v_{\varphi},z)\in\operatorname{End}({\mathcal{V}})[[z^{\pm 1}]]. The set S={vφ|φ}S=\{v_{\varphi}\,|\,\varphi\in{\mathcal{F}}\} generates 𝒱{\mathcal{V}}.

Proof.

We equip 𝒱{\mathcal{V}} with the representation of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{\mathbb{C}} from Lemma 3.10. To show that the point-like fields φ^(z)\hat{\varphi}(z) generate a Möbius vertex algebra, we invoke [RTT22, Thm. A.1] (see also [Kac98, Thm. 4.5]). To invoke his theorem, we need to verify that:

  1. 1.

    𝒱=n0ker(L0n){\mathcal{V}}=\bigoplus_{n\geq 0}\ker(L_{0}-n)

  2. 2.

    Ω\Omega is Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})-invariant

  3. 3.

    For every φ\varphi\in{\mathcal{F}}, φ^(z)Ω\hat{\varphi}(z)\Omega has a removable singularity at z=0z=0

  4. 4.

    For every φ\varphi\in{\mathcal{F}}, there exists a dφ0d_{\varphi}\in{\mathbb{Z}}_{\geq 0} such that

    [Lm,φ^(z)]=(zm+1ddz+(m+1)zmdφ)φ^(z)m=1,0,1[L_{m},\hat{\varphi}(z)]=(z^{m+1}\tfrac{d}{dz}+(m+1)z^{m}d_{\varphi})\hat{\varphi}(z)\qquad m=-1,0,1
  5. 5.

    For every φ,ψ\varphi,\psi\in{\mathcal{F}}, we have (zw)N[φ^(z),ψ^(w)]=0(z-w)^{N}[\hat{\varphi}(z),\hat{\psi}(w)]=0 for NN sufficiently large

  6. 6.

    𝒱=span{φ1(ej1)φk(ejk)Ω|k0,ji,φi}{\mathcal{V}}=\operatorname{span}\{\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega\,|\,k\geq 0,j_{i}\in{\mathbb{Z}},\varphi_{i}\in{\mathcal{F}}\}.

The first point follows from the fact that φ1(ej1)φk(ejk)Ω\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega is an eigenvector for L0L_{0} with eigenvalue ji-\sum j_{i} by the commutation relation of Lemma 3.10. The second point and fourth point also follow from Lemma 3.10 along with Equation (3.10). The fifth point holds by Lemma 3.12, and the sixth point is the definition of 𝒱{\mathcal{V}}.

We now argue the third point, that φ^(z)\hat{\varphi}(z) has a removable singularity at z=0z=0. The argument is the same as in [RTT22, Thm. 3.11]. Let φn=φ(en)\varphi_{n}=\varphi(e_{n}).333Note that φn\varphi_{n} is the nn-th mode of a single field φ\varphi and not the nn-th field. We use this notation only here and in the next paragraph.

We must show that φnΩ=0\varphi_{-n}\Omega=0 for nd1n\leq d-1. When n<0n<0 this identity holds by the spectrum condition which implies that ker(L0n)=0\ker(L_{0}-n)=0 for these nn. So we now consider n=0,,d1n=0,\ldots,d-1. From the L1L_{-1}-commutation relation of φ^\hat{\varphi} we have

φnΩ=1ndL1φn+1Ω.\varphi_{-n}\Omega=\tfrac{1}{n-d}L_{-1}\varphi_{-n+1}\Omega.

We repeatedly apply this identity, starting with n=0n=0, to obtain 0=φ0Ω==φd+1Ω0=\varphi_{0}\Omega=\cdots=\varphi_{-d+1}\Omega, as desired.

Thus by [RTT22, Thm. A.1] there exists a unique structure of Möbius vertex algebra on 𝒱{\mathcal{V}}, with the same LnL_{n}, such that for every φ\varphi\in{\mathcal{F}} with conformal dimension dd we have Y(φdΩ,z)=φ^(z)Y(\varphi_{-d}\Omega,z)=\hat{\varphi}(z). The vector φdΩ\varphi_{-d}\Omega is quasiprimary, as

L1φdΩ=[L1,φd]Ω=limz0[L1,Y(φdΩ,z)]Ω=limz0(z2ddz+2zd)Y(φdΩ,z)Ω=0.L_{1}\varphi_{-d}\Omega=[L_{1},\varphi_{-d}]\Omega=\lim_{z\to 0}[L_{1},Y(\varphi_{-d}\Omega,z)]\Omega=\lim_{z\to 0}(z^{2}\tfrac{d}{dz}+2zd)Y(\varphi_{-d}\Omega,z)\Omega=0.

By the sixth point, the set SS in the statement of the theorem generates 𝒱{\mathcal{V}}. This completes the proof of the existence statement.

For uniqueness, note that the set {limz0φ^(z)Ω}\{\lim_{z\to 0}\hat{\varphi}(z)\Omega\} generates any vertex algebra satisfying the statement of the theorem. The modes of the corresponding fields are determined by the fields φ(z)\varphi(z), and the modes of the remaining fields are then determined by the Borcherds product formula (2.5). The grading operator L0L_{0} is determined by the requirement that the conformal dimension of limz0φ^(z)Ω\lim_{z\to 0}\hat{\varphi}(z)\Omega matches the conformal dimension of φ\varphi. The operators L±1L_{\pm 1} are then determined by the commutation relations with the generating fields. We conclude that the Möbius vertex algebra constructed above is the unique such structure satisfying the requirements of the theorem. ∎

As a corollary of the proof of Theorem 3.13 we have that if φ\varphi\in{\mathcal{F}} is non-zero and has conformal dimension dd then

d=inf{n0|φ(en)Ω0},d=\inf\{n\in{\mathbb{Z}}_{\geq 0}\,|\,\varphi(e_{-n})\Omega\neq 0\}, (3.12)

and in particular this gives a proof that the conformal dimension of a Wightman field is uniquely determined.

We conclude this section with a canonical realization of the domain 𝒟{\mathcal{D}} of a Wightman CFT.

Proposition 3.14.

Let {\mathcal{F}} be a Möbius-covariant Wightman CFT on S1S^{1} with domain 𝒟{\mathcal{D}}, and let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} be the corresponding Möbius vertex algebra from Theorem 3.13. Equip 𝒟{\mathcal{D}} with the {\mathcal{F}}-strong topology and equip 𝒱^=n=0𝒱(n)\widehat{\mathcal{V}}=\prod_{n=0}^{\infty}{\mathcal{V}}(n) with the weak topology induced by 𝒱{\mathcal{V}}^{\prime}. Then the identity map id𝒱\operatorname{id}_{\mathcal{V}} extends to a (necessarily unique) injective continuous linear map ι:𝒟𝒱^\iota:{\mathcal{D}}\to\widehat{\mathcal{V}}.

Proof.

First, we claim that for any Φ𝒟\Phi\in{\mathcal{D}} there exists a unique sequence Φn𝒱(n)\Phi_{n}\in{\mathcal{V}}(n) such that Φ=Φn\Phi=\sum\Phi_{n}, converging in the {\mathcal{F}}-strong topology. We first consider existence. It suffices to establish existence for Φ=φ1(f1)φk(fk)Ω\Phi=\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega. Arguing as in the proof of Lemma 3.8, there exists a continuous map S:C((S1)k)𝒟^S:C^{\infty}((S^{1})^{k})\to\widehat{\mathcal{D}} such that

S(f1ππfk)=φ1(f1)φk(fk)ΩS(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k})=\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega

for all fjC(S1)f_{j}\in C^{\infty}(S^{1}), where 𝒟^\widehat{\mathcal{D}} is the completion of 𝒟{\mathcal{D}} in the {\mathcal{F}}-strong topology (see Appendix B) and

(f1ππfk)(z1,,zk)=f1(z1)fk(zk).(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k})(z_{1},\ldots,z_{k})=f_{1}(z_{1})\cdots f_{k}(z_{k}).

Let Pn:C((S1)k)C((S1)k)P_{n}:C^{\infty}((S^{1})^{k})\to C^{\infty}((S^{1})^{k}) be the projection onto the closed span of monomials z1n1zknkz_{1}^{n_{1}}\cdots z_{k}^{n_{k}} with n1++nk=nn_{1}+\cdots+n_{k}=-n (whose kernel is spanned by monomials with n1++nknn_{1}+\cdots+n_{k}\neq-n). When f1,,fk[z±1]f_{1},\ldots,f_{k}\in{\mathbb{C}}[z^{\pm 1}] we have

S(Pn(f1ππfk))𝒱(n;v1,,vk)S(P_{n}(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k}))\in{\mathcal{V}}(n;v_{1},\ldots,v_{k})

where vj𝒱v_{j}\in{\mathcal{V}} is the vector corresponding to φj\varphi_{j}. Since 𝒱(n;v1,,vk){\mathcal{V}}(n;v_{1},\ldots,v_{k}) has finite dimension, and is therefore complete [NB11, Thm. 4.10.3], the composed map SPnSP_{n} takes values in 𝒱(n;v1,,vk){\mathcal{V}}(n;v_{1},\ldots,v_{k}), and in particular in 𝒱(n){\mathcal{V}}(n). Thus if Φ=S(f1ππfk)\Phi=S(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k}) and we set Φn=SPn(f1ππfk)\Phi_{n}=SP_{n}(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k}), then Φ=Φn\Phi=\sum\Phi_{n} in (the natural extension of) the {\mathcal{F}}-strong topology on 𝒟^\widehat{\mathcal{D}}, because nPn(f1ππfk)=f1ππfk\sum_{n}P_{n}(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k})=f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k} in C((S1)k)C^{\infty}((S^{1})^{k}) and SS is continuous.

We now consider uniqueness of the sequence Φn\Phi_{n}. Suppose that Φn=0\sum\Phi_{n}=0 with Φn𝒱(n)\Phi_{n}\in{\mathcal{V}}(n) and the sum converging {\mathcal{F}}-strongly. Then any λ𝒱(m)𝒟\lambda\in{\mathcal{V}}(m)^{*}\cap{\mathcal{D}}_{\mathcal{F}}^{*} extends to 𝒟^\widehat{\mathcal{D}} by continuity (see Appendix B) and we have

0=λ(Φ)=λ(Φn)=λ(Φm).0=\lambda(\Phi)=\sum\lambda(\Phi_{n})=\lambda(\Phi_{m}).

As 𝒱(m)𝒟{\mathcal{V}}(m)^{*}\cap{\mathcal{D}}_{\mathcal{F}}^{*} separates points in 𝒱(m){\mathcal{V}}(m) by Lemma 2.11 we see Φm=0\Phi_{m}=0, and since mm was arbitrary this establishes the uniqueness portion of the claim.

We now define ι:𝒟𝒱^\iota:{\mathcal{D}}\to\widehat{\mathcal{V}} by ι(Φ)=(Φn)n0\iota(\Phi)=(\Phi_{n})_{n\geq 0}, where Φn𝒱(n)\Phi_{n}\in{\mathcal{V}}(n) is the unique sequence such that Φn=Φ\sum\Phi_{n}=\Phi with {\mathcal{F}}-strong convergence. This map is well-defined by the above claim and, by inspection, ι\iota is injective and restricts to the identity on 𝒱{\mathcal{V}}. It remains to check that ι\iota is continuous from the {\mathcal{F}}-strong topology to the weak topology on 𝒱^\widehat{\mathcal{V}} induced by 𝒱{\mathcal{V}}^{\prime}. By the universal property of the {\mathcal{F}}-strong topology, it suffices to check that λ(ιφ1(f1)φk(fk)Ω)\lambda(\iota\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega) depends continuously on the fjf_{j} for any λ𝒱(n)\lambda\in{\mathcal{V}}(n)^{*}. By the calculation above we have

λ(ιφ1(f1)φk(fk)Ω)=λ(SPn(f1ππfk)).\lambda(\iota\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega)=\lambda(SP_{n}(f_{1}\otimes_{\pi}\cdots\otimes_{\pi}f_{k})). (3.13)

We have seen that SPnSP_{n} is a continuous map with values in the finite-dimensional space 𝒱(n;v1,,vk){\mathcal{V}}(n;v_{1},\ldots,v_{k}), and λ|𝒱(n;v1,,vk)\lambda|_{{\mathcal{V}}(n;v_{1},\ldots,v_{k})} is evidently continuous. We conclude that (3.13) is continuous in the fjf_{j}, and so ι\iota is continuous as claimed. ∎

3.3 Equivalence of categories

We have constructions in Theorem 3.9 and Theorem 3.13 that produce Wightman CFTs from vertex algebras and vice versa. In this section we show that these constructions are inverse to each other, or more precisely we show that they induce an equivalence of categories. We now introduce the relevant categories.

A homomorphism g:𝒱𝒱~g:{\mathcal{V}}\to\tilde{\mathcal{V}} of Möbius vertex algebras is a linear map that intertwines the representations of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}), maps the vacuum vector to the vacuum vector, and intertwines the modes:

g(v(n)u)=g(v)(n)g(u)u,v𝒱.g(v_{(n)}u)=g(v)_{(n)}g(u)\qquad u,v\in{\mathcal{V}}.

Now suppose that 𝒱{\mathcal{V}} have 𝒱~\tilde{\mathcal{V}} are equipped with choices of generating sets of quasiprimary vectors SS and S~\tilde{S}, respectively. We say that gg is a morphism (𝒱,S)(𝒱~,S~)({\mathcal{V}},S)\to(\tilde{\mathcal{V}},\tilde{S}) if gg is a homomorphism of Möbius vertex algebras and g(S)S~g(S)\subset\tilde{S}. We write 𝖬𝖵𝖠+\mathsf{MVA}^{+} for the category of Möbius vertex algebras equipped with a choice of generating set of quasiprimary vectors.

If (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) and (~,𝒟~,U~,Ω~)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega}) are Möbius-covariant Wightman CFTs on S1S^{1}, then a morphism ~{\mathcal{F}}\to\tilde{\mathcal{F}} is a linear map g:𝒟𝒟~g:{\mathcal{D}}\to\tilde{\mathcal{D}} and a function g:~g_{*}:{\mathcal{F}}\to\tilde{\mathcal{F}} such that g(Ω)=Ω~g(\Omega)=\tilde{\Omega}, gg intertwines UU and U~\tilde{U}, and gφ(f)=(gφ)(f)gg\varphi(f)=(g_{*}\varphi)(f)g for all φ\varphi\in{\mathcal{F}} and fC(S1)f\in C^{\infty}(S^{1}). Note that gg_{*} is uniquely determined by gg. A straightforward calculation shows that a homomorphism gg is continuous when 𝒟{\mathcal{D}} and 𝒟~\tilde{\mathcal{D}} are respectively given the {\mathcal{F}}-strong and ~\tilde{\mathcal{F}}-strong topologies, and similarly for the {\mathcal{F}}-weak and ~\tilde{\mathcal{F}}-weak topologies. We write 𝖬𝖶\mathsf{MW} for the category of Möbius-covariant Wightman CFTs on S1S^{1}.

Lemma 3.15.

Let (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) and (~,𝒟~,U~,Ω~)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega}) be a pair of Möbius-covariant Wightman CFTs and let (g,g)(g,g_{*}) be a morphism ~{\mathcal{F}}\to\tilde{\mathcal{F}}. Let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} and 𝒱~𝒟~\tilde{\mathcal{V}}\subset\tilde{\mathcal{D}} be the Möbius vertex algebras constructed in Theorem 3.13, and let SS and S~\tilde{S} be the respective sets of generating vectors. Then g(𝒱)𝒱~g({\mathcal{V}})\subset\tilde{\mathcal{V}} and g|𝒱:(𝒱,S)(𝒱~,S~)g|_{\mathcal{V}}:({\mathcal{V}},S)\to(\tilde{\mathcal{V}},\tilde{S}) is a morphism in 𝖬𝖵𝖠+\mathsf{MVA}^{+}.

Proof.

By definition 𝒱{\mathcal{V}} is spanned by vectors of the form φ1(ej1)φk(ejk)Ω\varphi_{1}(e_{j_{1}})\cdots\varphi_{k}(e_{j_{k}})\Omega where φi\varphi_{i}\in{\mathcal{F}} and ej(z)=zje_{j}(z)=z^{j}. Since (g,g)(g,g_{*}) is a morphism we have

gφ1(ej1)φ(ejk)Ω=(gφ1)(ej1)(gφk)(ejk)Ω~𝒱~,g\varphi_{1}(e_{j_{1}})\cdots\varphi(e_{j_{k}})\Omega=(g_{*}\varphi_{1})(e_{j_{1}})\cdots(g_{*}\varphi_{k})(e_{j_{k}})\tilde{\Omega}\in\tilde{\mathcal{V}},

so g(𝒱)𝒱~g({\mathcal{V}})\subset\tilde{\mathcal{V}}.

We next check that gg intertwines the representations of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}). Let hddϑLie(Mo¨b)h\tfrac{d}{d\vartheta}\in\operatorname{Lie}(\operatorname{M\ddot{o}b}) and let γtMo¨b\gamma_{t}\in\operatorname{M\ddot{o}b} be the corresponding one-parameter group. We saw in the proof of Lemma 3.10 that the representations of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}) on 𝒟{\mathcal{D}} and 𝒟~\tilde{\mathcal{D}} are given by differentiating U(γt)U(\gamma_{t}), and so we have

gπ(hddϑ)v=gddt|t=0U(γt)v=ddt|t=0U~(γt)gv=π~(hddϑ)gvg\pi(h\tfrac{d}{d\vartheta})v=g\left.\frac{d}{dt}\right|_{t=0}U(\gamma_{t})v=\left.\frac{d}{dt}\right|_{t=0}\tilde{U}(\gamma_{t})gv=\tilde{\pi}(h\tfrac{d}{d\vartheta})gv

where we used that the derivatives are taken in the {\mathcal{F}}- and ~\tilde{\mathcal{F}}-weak topologies, and gg is continuous with respect to these topologies.

Now fix φ\varphi\in{\mathcal{F}} with conformal dimension dd. Let dd^{\prime} be the conformal dimension of gφg_{*}\varphi, and we begin by arguing d=dd=d^{\prime} provided gφ0g_{*}\varphi\not\equiv 0. By (3.12) we have

d=inf{n0|φ(en)Ω0},d=inf{n0|(gφ)(en)Ω0}.d=\inf\{n\in{\mathbb{Z}}_{\geq 0}\,|\,\varphi(e_{-n})\Omega\neq 0\},\qquad d^{\prime}=\inf\{n\in{\mathbb{Z}}_{\geq 0}\,|\,(g_{*}\varphi)(e_{-n})\Omega\neq 0\}.

As gφ(en)Ω=gφ(en)Ωg_{*}\varphi(e_{-n})\Omega=g\varphi(e_{-n})\Omega we have ddd\leq d^{\prime}, and we must show that gφ(ed)Ω0g\varphi(e_{-d})\Omega\neq 0. From the previous step we know that gLn=L~nggL_{n}=\tilde{L}_{n}g, where as usual Ln=π(ieinϑddϑ)L_{n}=\pi(-{\mathrm{i}}{\mathrm{e}}^{{\mathrm{i}}n\vartheta}\tfrac{d}{d\vartheta}) and similarly for L~n\tilde{L}_{n}. Thus for ndn\neq d we have

(gφ)(en)Ω~=gφ(en)Ω=1ndgL1φ(en+1)Ω=1ndL~1(gφ)(en+1)Ω~.(g_{*}\varphi)(e_{-n})\tilde{\Omega}=g\varphi(e_{-n})\Omega=\tfrac{1}{n-d}gL_{-1}\varphi(e_{-n+1})\Omega=\tfrac{1}{n-d}\tilde{L}_{-1}(g_{*}\varphi)(e_{-n+1})\tilde{\Omega}.

If (gφ)(ed)Ω~=0(g_{*}\varphi)(e_{-d})\tilde{\Omega}=0, we may apply the above relation repeatedly to n=d+1,d+2,n=d+1,d+2,\ldots to conclude that (gφ)(en)Ω~=0(g_{*}\varphi)(e_{-n})\tilde{\Omega}=0 for all n0n\in{\mathbb{Z}}_{\geq 0}. But then we would have (gφ)(ed)Ω~=0(g_{*}\varphi)(e_{-d^{\prime}})\tilde{\Omega}=0, a contradiction. We conclude that d=dd^{\prime}=d, which is to say that φ\varphi and gφg_{*}\varphi have the same conformal dimension provided gφ0g_{*}\varphi\not\equiv 0.

Next observe that Y(gv,z)=(gφ^)(z)Y(gv,z)=(\widehat{g_{*}\varphi})(z), or equivalently (gv)(n)=(gφ)(end+1)(gv)_{(n)}=(g_{*}\varphi)(e_{n-d+1}). We therefore have

gv(n)=gφ(end+1)=(gφ)(end+1)g=(gv)(n)g.gv_{(n)}=g\varphi(e_{n-d+1})=(g_{*}\varphi)(e_{n-d+1})g=(gv)_{(n)}g.

This means that gg intertwines the actions of modes of vectors vv corresponding to φ\varphi\in{\mathcal{F}}, and since such vectors generate 𝒱{\mathcal{V}} we can conclude that gg intertwines the actions of modes v(n)v_{(n)} for all v𝒱v\in{\mathcal{V}}. Moreover the identity gφ(ed)Ω=(gφ)(ed)Ω~g\varphi(e_{-d})\Omega=(g_{*}\varphi)(e_{-d})\tilde{\Omega} implies that gSS~gS\subset\tilde{S}. ∎

Lemma 3.16.

Let 𝒱{\mathcal{V}} and 𝒱~\tilde{\mathcal{V}} be Möbius vertex algebras with generating sets SS and S~\tilde{S}, respectively. Let g:(𝒱,S)(𝒱~,S~)g:({\mathcal{V}},S)\to(\tilde{\mathcal{V}},\tilde{S}) be a morphism in 𝖬𝖵𝖠+\mathsf{MVA}^{+}. Let (𝒟,,U,Ω)({\mathcal{D}},{\mathcal{F}},U,\Omega) and (𝒟~,,U~,Ω~)(\tilde{\mathcal{D}},{\mathcal{F}},\tilde{U},\tilde{\Omega}) be the Möbius-covariant Wightman CFTs constructed in Theorem 3.9. Then there is a unique morphism (h,h):~(h,h_{*}):{\mathcal{F}}\to\tilde{\mathcal{F}} such that h|𝒱=gh|_{\mathcal{V}}=g.

Proof.

For vSv\in S, we write φv:=Y(v,)\varphi_{v}:=Y(v,\cdot) for the corresponding Wightman field in {\mathcal{F}}, and similarly for v~S~\tilde{v}\in\tilde{S} we write φ~v~:=Y~(v~,)\tilde{\varphi}_{\tilde{v}}:=\tilde{Y}(\tilde{v},\cdot) for the Wightman field in ~\tilde{\mathcal{F}}. For vSv\in S, we define hφv=φ~gv~h_{*}\varphi_{v}=\tilde{\varphi}_{gv}\in\tilde{\mathcal{F}}. Since gg is a morphism of vertex algebras we have for all φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and all f1,,fk[z±1]f_{1},\ldots,f_{k}\in{\mathbb{C}}[z^{\pm 1}]

gφ1(f1)φk(fk)Ω=(hφ1)(f1)(hφk)(fk)Ω~.g\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega=(h_{*}\varphi_{1})(f_{1})\cdots(h_{*}\varphi_{k})(f_{k})\tilde{\Omega}.

Since morphisms of Wightman CFTs are continuous for the {\mathcal{F}}- and ~\tilde{\mathcal{F}}-weak topologies, we can see from the above formula that a morphism (h,h)(h,h_{*}) as in the statement of the lemma is necessarily unique.

Since gg intertwines the actions of L0L_{0} and L~0\tilde{L}_{0}, the adjoint operator gg^{*} maps 𝒱~\tilde{\mathcal{V}}^{\prime} into 𝒱{\mathcal{V}}^{\prime}. As 𝒱𝒟{\mathcal{V}}^{\prime}\subset{\mathcal{D}}_{\mathcal{F}}^{*} and 𝒱~𝒟~~\tilde{\mathcal{V}}^{\prime}\subset\tilde{\mathcal{D}}_{\tilde{\mathcal{F}}}^{*} (by Lemma 3.4), and 𝒱~\tilde{\mathcal{V}}^{\prime} separates points in 𝒟~\tilde{\mathcal{D}}, it follows that the closure of the graph of g:𝒱𝒱~g:{\mathcal{V}}\to\tilde{\mathcal{V}} in 𝒟×𝒟~{\mathcal{D}}\times\tilde{\mathcal{D}} is again the graph of a densely defined linear map h:𝒟𝒟~h:{\mathcal{D}}\to\tilde{\mathcal{D}}. If f1,,fkC(S1)f_{1},\ldots,f_{k}\in C^{\infty}(S^{1}), we may approximate each fjf_{j} by Laurent polynomials fj,nf_{j,n} to obtain

hφ1(f1,n)φk(fk,n)Ω=(hφ1)(f1,n)(hφk)(fk,n)Ω(hφ1)(f1)(hφk)(fk)Ω.h\varphi_{1}(f_{1,n})\cdots\varphi_{k}(f_{k,n})\Omega=(h_{*}\varphi_{1})(f_{1,n})\cdots(h_{*}\varphi_{k})(f_{k,n})\Omega\to(h_{*}\varphi_{1})(f_{1})\cdots(h_{*}\varphi_{k})(f_{k})\Omega.

Hence hh is defined on all of 𝒟{\mathcal{D}} and hφ(f)=(hφ)(f)hh\varphi(f)=(h_{*}\varphi)(f)h for all fC(S1)f\in C^{\infty}(S^{1}) and φ\varphi\in{\mathcal{F}}. It follows immediately that hh also intertwines the representations UU and U~\tilde{U}, and we have shown that (h,h)(h,h_{*}) is a morphism ~{\mathcal{F}}\to\tilde{\mathcal{F}}. ∎

Lemmas 3.15 and 3.16 upgrade the constructions of Theorem 3.13 and 3.9 to a pair of functors F:𝖬𝖶𝖬𝖵𝖠+F:\mathsf{MW}\to\mathsf{MVA}^{+} and G:𝖬𝖵𝖠+𝖬𝖶G:\mathsf{MVA}^{+}\to\mathsf{MW}. In showing that these are an equivalence of categories, it will be helpful to note that if {\mathcal{F}} is a Wightman CFT with domain 𝒟{\mathcal{D}}, then the vertex algebra 𝒱:=F(){\mathcal{V}}:=F({\mathcal{F}}) is a subspace 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}}. Conversely, if 𝒱𝖬𝖵𝖠+{\mathcal{V}}\in\mathsf{MVA}^{+} and 𝒟{\mathcal{D}} is the domain of the Wightman CFT G(𝒱)G({\mathcal{V}}), then 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}}.

Lemma 3.17.

We have the following.

  1. i)

    Let (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) be a Möbius-covariant Wightman CFT, let 𝒱=F(){\mathcal{V}}=F({\mathcal{F}}) with 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}}. Let (~,𝒟~,U~,Ω~)=G(𝒱)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega})=G({\mathcal{V}}) with 𝒱𝒟~{\mathcal{V}}\subset\tilde{\mathcal{D}}. Then there is a unique isomorphism (g,g):~(g,g_{*}):{\mathcal{F}}\to\tilde{\mathcal{F}} such that g|𝒱=idg|_{\mathcal{V}}=\mathrm{id}.

  2. ii)

    Let 𝒱{\mathcal{V}} be a Möbius vertex algebra, let (,𝒟,U,Ω)=G(𝒱)({\mathcal{F}},{\mathcal{D}},U,\Omega)=G({\mathcal{V}}) be the corresponding Möbius-covariant Wightman CFT with 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}}. Let 𝒱~𝒟\tilde{\mathcal{V}}\subset{\mathcal{D}} be the Möbius vertex algebra F()F({\mathcal{F}}). Then 𝒱~=𝒱\tilde{\mathcal{V}}={\mathcal{V}} as Möbius vertex algebras.

Proof.

We first consider (i). Uniqueness of such an isomorphism follows from the fact that an isomorphism g:𝒟𝒟~g:{\mathcal{D}}\to\tilde{\mathcal{D}} is {\mathcal{F}}-strong continuous and 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} is {\mathcal{F}}-strong dense. We now consider existence. By construction there is a canonical bijection ~{\mathcal{F}}\to\tilde{\mathcal{F}} which we denote by φφ~\varphi\mapsto\tilde{\varphi}. We must verify that there exists a corresponding bijection 𝒟𝒟~{\mathcal{D}}\to\tilde{\mathcal{D}}. We have 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} and 𝒱𝒟~{\mathcal{V}}\subset\tilde{\mathcal{D}}. By construction we have 𝒟~𝒱^\tilde{\mathcal{D}}\subset\widehat{\mathcal{V}}, and Proposition 3.14 provides a map ι:𝒟𝒱^\iota:{\mathcal{D}}\hookrightarrow\widehat{\mathcal{V}}. We have ιφ1(f1)φk(fk)Ω=φ~1(f1)φ~k(fk)Ω\iota\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega=\tilde{\varphi}_{1}(f_{1})\cdots\tilde{\varphi}_{k}(f_{k})\Omega when fj[z±1]f_{j}\in{\mathbb{C}}[z^{\pm 1}], and since both sides are continuous in the functions fjf_{j} this extends to all fjC(S1)f_{j}\in C^{\infty}(S^{1}). We conclude that ι\iota maps 𝒟{\mathcal{D}} into 𝒟~\tilde{\mathcal{D}} and furnishes the necessary bijection. Part (ii) is immediate from the construction. ∎

The isomorphisms from Lemma 3.17 are natural in {\mathcal{F}} and 𝒱{\mathcal{V}} respectively, and thus we have proven the following.

Theorem 3.18.

Let 𝖬𝖶\mathsf{MW} be the category of (not-necessarily-unitary) Möbius-covariant Wightman CFTs and let 𝖬𝖵𝖠+\mathsf{MVA}^{+} be the category of Möbius vertex algebras equipped with a generating family of quasiprimary vectors. Let F:𝖬𝖶𝖬𝖵𝖠+F:\mathsf{MW}\to\mathsf{MVA}^{+} be the functor constructed on objects in Theorem 3.13 and on morphisms in Lemma 3.15. Let G:𝖬𝖵𝖠+𝖬𝖶G:\mathsf{MVA}^{+}\to\mathsf{MW} be the functor constructed on objects in Theorem 3.9 and on morphisms in Lemma 3.16. Then FF and GG, along with the isomorphisms of Lemma 3.17, are an equivalence of categories between 𝖬𝖶\mathsf{MW} and 𝖬𝖵𝖠+\mathsf{MVA}^{+}.

4 Invariant forms and unitary theories

In this section we show that the correspondence between Wightman CFTs on S1S^{1} and Möbius vertex algebras constructed in Section 3 is compatible with invariant bilinear forms. The definition of an invariant bilinear form for a Möbius vertex algebra is standard (see [FHL93, §5.2] and [Li94]).

Definition 4.1.

An invariant bilinear form (,)(\cdot,\cdot) on a Möbius vertex algebra 𝒱{\mathcal{V}} is a bilinear form such that

(Y(v,z)u1,u2)=(u1,Y(ezL1(z2)L0v,z1)u2)(Y(v,z)u_{1},u_{2})=(u_{1},Y({\mathrm{e}}^{zL_{1}}(-z^{-2})^{L_{0}}v,z^{-1})u_{2}) (4.1)

and

(Lnu1,u2)=(u1,Lnu2)(L_{n}u_{1},u_{2})=(u_{1},L_{-n}u_{2}) (4.2)

for all v,u1,u2𝒱v,u_{1},u_{2}\in{\mathcal{V}}.

It can be convenient to introduce notation for the opposite vertex operator

Y𝗈(v,z)=Y(ezL1(z2)L0v,z1),Y^{\mathsf{o}}(v,z)=Y({\mathrm{e}}^{zL_{1}}(-z^{-2})^{L_{0}}v,z^{-1}), (4.3)

and in this notation the invariance condition becomes

(Y(v,z)u1,u2)=(u1,Y𝗈(v,z)u2).(Y(v,z)u_{1},u_{2})=(u_{1},Y^{\mathsf{o}}(v,z)u_{2}).

The map LnLnL_{n}\mapsto-L_{-n} extends linearly to a Lie algebra automorphism of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{{\mathbb{C}}} which leaves Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}) invariant. Let dα:Lie(Mo¨b)Lie(Mo¨b)d\alpha:\operatorname{Lie}(\operatorname{M\ddot{o}b})\to\operatorname{Lie}(\operatorname{M\ddot{o}b}) be this restriction. In this notation, the compatibility condition (4.2) between the invariant bilinear form and the representation π\pi of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}) on 𝒱{\mathcal{V}} becomes

(π(fddϑ)u1,u2)=(u1,π(dα(fddϑ))u2).(\pi(f\tfrac{d}{d\vartheta})u_{1},u_{2})=-(u_{1},\pi(d\alpha(f\tfrac{d}{d\vartheta}))u_{2}).

In order to formulate the correct notion of invariant bilinear form for a Wightman CFT, we must integrate dαd\alpha to an automorphism α\alpha of Mo¨b\operatorname{M\ddot{o}b}. It is straightforward to check that α\alpha is given by

(αγ)(z)=1/γ(1z).(\alpha\gamma)(z)=1/\gamma(\tfrac{1}{z}).

Indeed, at the level of matrices α\alpha is given on (abb¯a¯)SU(1,1)\begin{pmatrix}a&b\\ \overline{b}&\overline{a}\end{pmatrix}\in\mathrm{SU}(1,1) (with |a|2|b|2=1\left|a\right|^{2}-\left|b\right|^{2}=1) by complex conjugation

α(abb¯a¯)=(a¯b¯ba)\alpha\begin{pmatrix}a&b\\ \overline{b}&\overline{a}\end{pmatrix}=\begin{pmatrix}\overline{a}&\overline{b}\\ b&a\end{pmatrix}

and dαd\alpha is given on (icdd¯ic)𝔰𝔲(1,1)\begin{pmatrix}{\mathrm{i}}c&d\\ \overline{d}&-{\mathrm{i}}c\end{pmatrix}\in\mathfrak{su}(1,1) (with cc\in{\mathbb{R}}) by complex conjugation as well

dα(icdd¯ic)=(icd¯dic).d\alpha\begin{pmatrix}{\mathrm{i}}c&d\\ \overline{d}&-{\mathrm{i}}c\end{pmatrix}=\begin{pmatrix}-{\mathrm{i}}c&\overline{d}\\ d&{\mathrm{i}}c\end{pmatrix}.

In particular we have

exp(dα(fddϑ))=α(exp(fddϑ)).\exp(d\alpha(f\tfrac{d}{d\vartheta}))=\alpha(\exp(f\tfrac{d}{d\vartheta})). (4.4)

We thus have the following notion of invariant bilinear form for a Wightman CFT.

Definition 4.2.

Let (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) be a Möbius-covariant Wightman CFT on S1S^{1}. A jointly {\mathcal{F}}-strong continuous bilinear form (,)(\,\cdot\,,\,\cdot\,) on 𝒟{\mathcal{D}} is called an invariant bilinear form if

(φ(f)Φ,Ψ)=(Φ,(1)dφφ(f1z)Ψ)(\varphi(f)\Phi,\Psi)=(\Phi,(-1)^{d_{\varphi}}\varphi(f\circ\tfrac{1}{z})\Psi) (4.5)

for all φ\varphi\in{\mathcal{F}} (with conformal dimension dφd_{\varphi}), all fC(S1)f\in C^{\infty}(S^{1}), and all Φ,Ψ𝒟\Phi,\Psi\in{\mathcal{D}}, and moreover

(U(γ)Φ,U(α(γ))Ψ)=(Φ,Ψ)(U(\gamma)\Phi,U(\alpha(\gamma))\Psi)=(\Phi,\Psi) (4.6)

for all γMo¨b\gamma\in\operatorname{M\ddot{o}b} and Φ,Ψ𝒟\Phi,\Psi\in{\mathcal{D}}.

As in the context of vertex algebras, we can introduce the notion of opposite field

φ𝗈(f):=(1)dφφ(f1z)\varphi^{\mathsf{o}}(f):=(-1)^{d_{\varphi}}\varphi(f\circ\tfrac{1}{z})

and the invariance condition (4.5) then becomes

(φ(f)Φ,Ψ)=(Φ,φ𝗈(f)Ψ).(\varphi(f)\Phi,\Psi)=(\Phi,\varphi^{\mathsf{o}}(f)\Psi).
Theorem 4.3 (Correspondence between invariant bilinear forms).

Let (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) be a Möbius-covariant Wightman CFT on S1S^{1} and let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} be the corresponding Möbius vertex algebra. Then

  1. i)

    Every invariant bilinear form for the Wightman CFT 𝒟{\mathcal{D}} restricts to an invariant bilinear form for the vertex algebra 𝒱{\mathcal{V}}.

  2. ii)

    Every invariant bilinear form for the vertex algebra 𝒱{\mathcal{V}} extends uniquely to an invariant bilinear form for the Wightman CFT on 𝒟{\mathcal{D}}.

If an invariant form on 𝒱{\mathcal{V}} is nondegenerate, then so is the extension to 𝒟{\mathcal{D}}. Conversely, if an invariant form on 𝒟{\mathcal{D}} is nondegenerate then so is the restriction to 𝒱{\mathcal{V}}.

Proof.

First suppose that 𝒟{\mathcal{D}} is equipped with an invariant bilinear form (,)(\,\cdot\,,\,\cdot\,). Let XLie(Mo¨b)X\in\operatorname{Lie}(\operatorname{M\ddot{o}b}), let γt=exp(tX)Mo¨b\gamma_{t}=\exp(tX)\in\operatorname{M\ddot{o}b}, and let

ρt=α(exp(tX))=exp(tdα(X)).\rho_{t}=\alpha(\exp(tX))=\exp(td\alpha(X)).

For u1,u2𝒱u_{1},u_{2}\in{\mathcal{V}} we have

(U(γt)u1,u2)=(u1,U(ρt)u2).(U(\gamma_{t})u_{1},u_{2})=(u_{1},U(\rho_{-t})u_{2}).

Differentiating and evaluating at t=0t=0 (as in the proof of Lemma 3.10) yields

(π(X)u1,u2)=(u1,π(dα(X))u2),(\pi(X)u_{1},u_{2})=-(u_{1},\pi(d\alpha(X))u_{2}),

as required. Now let S𝒱S\subset{\mathcal{V}} be the set of quasiprimary generators corresponding to {\mathcal{F}}. For vSv\in S we have

(Y(v,f)u1,u2)=(1)dv(u1,Y(v,f1z)u2)(Y(v,f)u_{1},u_{2})=(-1)^{d_{v}}(u_{1},Y(v,f\circ\tfrac{1}{z})u_{2})

and in particular at the level of modes vnEnd(𝒱)v_{n}\in\mathrm{End}({\mathcal{V}}) we have

(vnu1,u2)=(1)dv(u1,vnu2).(v_{n}u_{1},u_{2})=(-1)^{d_{v}}(u_{1},v_{-n}u_{2}).

Hence for vSv\in S we have

(Y(v,z)u1,u2)=(u1,Y𝗈(v,z)u2).(Y(v,z)u_{1},u_{2})=(u_{1},Y^{\mathsf{o}}(v,z)u_{2}).

This extends to all v𝒱v\in{\mathcal{V}} by Lemma 4.4 below, and we have established (i).

Now conversely suppose that 𝒱{\mathcal{V}} is equipped with an invariant bilinear form which we denote (,)𝒱(\,\cdot\,,\,\cdot\,)_{{\mathcal{V}}}. Note that a {\mathcal{F}}-strongly continuous extension of such a form on 𝒱{\mathcal{V}} to a bilinear form on 𝒟{\mathcal{D}} is unique, and so we must only show existence. Recall from Proposition 3.14 that 𝒟{\mathcal{D}} comes naturally embedded in 𝒱^=n=0𝒱(n)\widehat{\mathcal{V}}=\prod_{n=0}^{\infty}{\mathcal{V}}(n). The bilinear form on 𝒱{\mathcal{V}} naturally extends to a pairing of 𝒱{\mathcal{V}} and 𝒱^\widehat{\mathcal{V}}. First, we claim that for φ1,,φk\varphi_{1},\ldots,\varphi_{k}, ψ1,,ψ\psi_{1},\ldots,\psi_{\ell}\in{\mathcal{F}} and f1,,fk,g1,,gC(S1)f_{1},\ldots,f_{k},g_{1},\ldots,g_{\ell}\in C^{\infty}(S^{1}) we have

(ψ𝗈(g)ψ𝗈(g1)φ1(f1)φk(fk)Ω,Ω)𝒱^,𝒱=(Ω,φ𝗈(fk)φ𝗈(f1)ψ(g1)ψ(g)Ω)𝒱,𝒱^.(\psi^{\mathsf{o}}(g_{\ell})\cdots\psi^{\mathsf{o}}(g_{1})\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega,\Omega)_{\widehat{\mathcal{V}},{\mathcal{V}}}=(\Omega,\varphi^{\mathsf{o}}(f_{k})\cdots\varphi^{\mathsf{o}}(f_{1})\psi(g_{1})\cdots\psi(g_{\ell})\Omega)_{{\mathcal{V}},\widehat{\mathcal{V}}}. (4.7)

Indeed these agree when fi,gj[z±1]f_{i},g_{j}\in{\mathbb{C}}[z^{\pm 1}] since the form is invariant for 𝒱{\mathcal{V}}, and this identity extends to all smooth functions by continuity.

With this in mind, we wish to define a bilinear form on 𝒟{\mathcal{D}} by extending linearly the prescription

(φ1(f1)φk(fk)Ω,ψ1(g1)ψ(g)Ω)𝒟:=(ψ𝗈(g)ψ𝗈(g1)φ1(f1)φk(fk)Ω,Ω)𝒱^,𝒱,(\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega,\psi_{1}(g_{1})\cdots\psi_{\ell}(g_{\ell})\Omega)_{\mathcal{D}}:=(\psi^{\mathsf{o}}(g_{\ell})\cdots\psi^{\mathsf{o}}(g_{1})\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega,\Omega)_{\widehat{\mathcal{V}},{\mathcal{V}}}, (4.8)

but we must first check that this is well-defined. Suppose that for some collection of Wightman fields φi,j\varphi_{i,j}\in{\mathcal{F}} and smearing functions fi,jC(S1)f_{i,j}\in C^{\infty}(S^{1}) we have

0=iφ1,j(f1,j)φkj,j(fkj,j)Ω=0.0=\sum_{i}\varphi_{1,j}(f_{1,j})\cdots\varphi_{k_{j},j}(f_{k_{j},j})\Omega=0.

Then for all ψ1,,ψ\psi_{1},\ldots,\psi_{\ell}\in{\mathcal{F}} and g1,,gC(S1)g_{1},\ldots,g_{\ell}\in C^{\infty}(S^{1})

0=jψ𝗈(g)ψ𝗈(g1)φ1,j(f1,j)φkj,j(fkj,j)Ω,0=\sum_{j}\psi^{\mathsf{o}}(g_{\ell})\cdots\psi^{\mathsf{o}}(g_{1})\varphi_{1,j}(f_{1,j})\cdots\varphi_{k_{j},j}(f_{k_{j},j})\Omega,

and thus

0=j(ψ𝗈(g)ψ𝗈(g1)φ1,j(f1,j)φkj,j(fkj,j)Ω,Ω)𝒱^,𝒱.0=\sum_{j}(\psi^{\mathsf{o}}(g_{\ell})\cdots\psi^{\mathsf{o}}(g_{1})\varphi_{1,j}(f_{1,j})\cdots\varphi_{k_{j},j}(f_{k_{j},j})\Omega,\Omega)_{\widehat{\mathcal{V}},{\mathcal{V}}}.

This shows that the prescription (4.8) is well-defined in the first input. We may repeat the above argument (invoking (4.7)) to show that it is also well-defined in the second input, and we conclude that (4.8) extends to a well-defined bilinear form. As (4.8) is continuous in the functions fjf_{j} and gjg_{j}, the bilinear form on 𝒟{\mathcal{D}} is jointly {\mathcal{F}}-strong continuous, as required.

Finally, we need to check that (,)𝒟,𝒟(\,\cdot\,,\cdot\,)_{{\mathcal{D}},{\mathcal{D}}} is compatible with the representation UU of Mo¨b\operatorname{M\ddot{o}b}. Let XLie(Mo¨b)X\in\operatorname{Lie}(\operatorname{M\ddot{o}b}), let γt=exp(tX)\gamma_{t}=\exp(tX) and recall that α(γt)=exp(tdα(X))\alpha(\gamma_{t})=\exp(td\alpha(X)). From the proof of Lemma 3.6 we have for Φ𝒟\Phi\in{\mathcal{D}}

ddtU(γt)Φ=ddsU(γt+s)Φ|s=0=ddsU(γs)U(γt)Φ|s=0=π(X)U(γt)Φ,\frac{d}{dt}U(\gamma_{t})\Phi=\left.\frac{d}{ds}U(\gamma_{t+s})\Phi\right|_{s=0}=\left.\frac{d}{ds}U(\gamma_{s})U(\gamma_{t})\Phi\right|_{s=0}=\pi(X)U(\gamma_{t})\Phi,

with the derivative taken in the {\mathcal{F}}-strong topology on 𝒟{\mathcal{D}}. Similarly

ddtU(α(γt))Ψ=π(dα(X))U(α(γt))Ψ.\frac{d}{dt}U(\alpha(\gamma_{t}))\Psi=\pi(d\alpha(X))U(\alpha(\gamma_{t}))\Psi.

Hence by the joint continuity of the bilinear form we have

ddt(U(γt)Φ,U(α(γt))Ψ)𝒟\displaystyle\frac{d}{dt}(U(\gamma_{t})\Phi,U(\alpha(\gamma_{t}))\Psi)_{\mathcal{D}}
=(π(X)U(γt)Φ,U(α(γt))Ψ)𝒟+(U(γt)Φ,π(dα(X))U(α(γt))Ψ)𝒟\displaystyle=(\pi(X)U(\gamma_{t})\Phi,U(\alpha(\gamma_{-t}))\Psi)_{\mathcal{D}}+(U(\gamma_{t})\Phi,\pi(d\alpha(X))U(\alpha(\gamma_{t}))\Psi)_{\mathcal{D}}
=0.\displaystyle=0.

In the last equality we used the fact that (π(X)u1,u2)=(u1,π(dα(X))u2)(\pi(X)u_{1},u_{2})=-(u_{1},\pi(d\alpha(X))u_{2}) for uj𝒱u_{j}\in{\mathcal{V}}, which extends to vectors in 𝒟{\mathcal{D}} by the {\mathcal{F}}-strong continuity of π(X)\pi(X) and π(dα(X))\pi(d\alpha(X)). Hence the above expression is independent of tt, and as the exponential map Lie(Mo¨b)Mo¨b\operatorname{Lie}(\operatorname{M\ddot{o}b})\to\operatorname{M\ddot{o}b} is surjective we then have

(Φ,Ψ)𝒟=(U(γ)Φ,U(α(γ))Ψ)𝒟(\Phi,\Psi)_{{\mathcal{D}}}=(U(\gamma)\Phi,U(\alpha(\gamma))\Psi)_{\mathcal{D}}

for all γMo¨b\gamma\in\operatorname{M\ddot{o}b} and Φ,Ψ𝒟\Phi,\Psi\in{\mathcal{D}}, as required.

Finally, we address nondegeneracy. Recall that 𝒟{\mathcal{D}} embeds naturally in 𝒱^\widehat{\mathcal{V}} and that the bilinear form on 𝒟{\mathcal{D}} is compatible with the pairing of 𝒱{\mathcal{V}} and 𝒱^\widehat{\mathcal{V}}. If the bilinear form on 𝒱{\mathcal{V}} is nondegenerate and Φ=(Φn)n0𝒟\Phi=(\Phi_{n})_{n\geq 0}\in{\mathcal{D}} (with Φn𝒱(n)\Phi_{n}\in{\mathcal{V}}(n)) is non-zero, then Φn0\Phi_{n}\neq 0 for some nn, and thus there exists v𝒱(n)v\in{\mathcal{V}}(n) such that

(Φ,v)=(Φn,v)0.(\Phi,v)=(\Phi_{n},v)\neq 0.

A similar argument shows that the right-kernel of the form is zero, and so the form on 𝒟{\mathcal{D}} is nondegenerate.

Conversely, assume that the form on 𝒟{\mathcal{D}} is nongenerate. By Möbius invariance, its restriction to 𝒱{\mathcal{V}} is the direct sum (v,u)=n(v,u)(v,u)=\sum_{n}(v,u). Let v𝒱v\in{\mathcal{V}} with v0v\neq 0. Then there exists Φ=(Φn)n0𝒟\Phi=(\Phi_{n})_{n\geq 0}\in{\mathcal{D}} with

0(v,Φ)=n(vn,Φn).0\neq(v,\Phi)=\sum_{n}(v_{n},\Phi_{n}).

Hence there must be some nn such that (vn,Φn)0(v_{n},\Phi_{n})\neq 0, and so the left-kernel of the form on 𝒱{\mathcal{V}} is zero. A similar argument shows that right-kernel is zero as well. ∎

We used the following fact in the proof of Theorem 4.3.

Lemma 4.4.

Let 𝒱{\mathcal{V}} be a Möbius vertex algebra equipped with a bilinear form (,)(\,\cdot\,,\,\cdot\,), and let S𝒱S\subset{\mathcal{V}} be a set of vectors that generate 𝒱{\mathcal{V}}. Suppose that the invariance condition

(Y(v,z)u1,u2)=(u1,Y𝗈(v,z)u2)(Y(v,z)u_{1},u_{2})=(u_{1},Y^{\mathsf{o}}(v,z)u_{2})

holds for vSv\in S and u1,u2𝒱u_{1},u_{2}\in{\mathcal{V}}, and also that

(Lnu1,u2)=(u1,Lnu2)(L_{n}u_{1},u_{2})=(u_{1},L_{-n}u_{2})

for all u1,u2𝒱u_{1},u_{2}\in{\mathcal{V}}. Then the invariance condition holds for all v𝒱v\in{\mathcal{V}} (i.e.​ the form is an invariant bilinear form for 𝒱{\mathcal{V}}).

Proof.

There is a (generalized) 𝒱{\mathcal{V}}-module structure on the restricted dual 𝒱{\mathcal{V}}^{\prime} whose state-field correspondence Y(v,z)Y^{\prime}(v,z) is characterized by

(Y𝗈(v,z)u,u)𝒱,𝒱=(u,Y(v,z)u)𝒱,𝒱(Y^{\mathsf{o}}(v,z)u,u^{\prime})_{{\mathcal{V}},{\mathcal{V}}^{\prime}}=(u,Y^{\prime}(v,z)u^{\prime})_{{\mathcal{V}},{\mathcal{V}}^{\prime}}

for all u,v𝒱u,v\in{\mathcal{V}} and u𝒱u^{\prime}\in{\mathcal{V}}^{\prime}. This contragredient module structure was first studied in [FHL93, §5.2] and described further in our context with infinite-dimensional weight spaces in the paragraphs following [HLZ14, Lem.​ 2.22]. If f:𝒱𝒱f:{\mathcal{V}}\to{\mathcal{V}}^{\prime} is the map f(u)=(u,)f(u)=(u,\,\cdot\,), then our hypothesis implies that fY(v,z)=Y(v,z)ffY(v,z)=Y^{\prime}(v,z)f for all vSv\in S, or at the level of modes fv(n)=v(n)ffv_{(n)}=v^{\prime}_{(n)}f for all vSv\in S and nn\in{\mathbb{Z}}. This intertwining condition extends to all v𝒱v\in{\mathcal{V}} by the Borcherds product formula (for 𝒱{\mathcal{V}} and for 𝒱{\mathcal{V}}^{\prime}), and we conclude that the bilinear form is invariant. ∎

It was shown in [FHL93, Prop. 5.3.6] that every nondegenerate invariant bilinear form on a vertex operator algebra is symmetric. Later it was observed in [Li94, Prop. 2.6] that the proof does not use the hypothesis of nondegeneracy, and further examination of the proof in [FHL93] shows that the proof also goes through for Möbius vertex algebras as defined in this article (that is, allowing for infinite-dimensional L0L_{0}-weight spaces and only using Möbius symmetry rather than Virasoro). In light of Theorem 4.3, we have the same result for Wightman CFTs.

Corollary 4.5.

Every invariant bilinear form on a Wightman CFT is symmetric.

We now turn our attention to unitary theories, and more generally invariant sesquilinear forms (which we call involutive structures). In order to do this we will need to introduce antilinear homomorphisms of Möbius vertex algebras and Möbius-covariant Wightman CFTs. Let 𝒱{\mathcal{V}} and 𝒱~\tilde{\mathcal{V}} be Möbius vertex algebras, with vacuum vectors Ω\Omega and Ω~\tilde{\Omega} and representations LnL_{n} and L~n\tilde{L}_{n} of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b})_{{\mathbb{C}}}, respectively. Then an antilinear map g:𝒱𝒱~g:{\mathcal{V}}\to\tilde{\mathcal{V}} is called a homomorphism if g(Ω)=Ω~g(\Omega)=\tilde{\Omega} and

gv(m)=(gv)(m)g, and gLn=L~nggv_{(m)}=(gv)_{(m)}g,\qquad\mbox{ and }\qquad gL_{n}=\tilde{L}_{n}g

for all v𝒱v\in{\mathcal{V}}, mm\in{\mathbb{Z}} and n=1,0,1n=-1,0,1.

On the Wightman side, if (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) and (~,𝒟~,U~,Ω~)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega}) are Möbius-covariant Wightman CFTs, an antilinear homomorphism ~{\mathcal{F}}\to\tilde{\mathcal{F}} is an antilinear map g:𝒟𝒟~g:{\mathcal{D}}\to\tilde{\mathcal{D}} and a function g:~g_{*}:{\mathcal{F}}\to\tilde{\mathcal{F}} such that g(Ω)=Ω~g(\Omega)=\tilde{\Omega} and

gφ(f)=(gφ)(f¯1z)g, and gU(γ)=U~(α(γ))gg\varphi(f)=(g_{*}\varphi)(\overline{f}\circ\tfrac{1}{z})g,\qquad\mbox{ and }\qquad gU(\gamma)=\tilde{U}(\alpha(\gamma))g

for all φ\varphi\in{\mathcal{F}}, fC(S1)f\in C^{\infty}(S^{1}) and γMo¨b\gamma\in\operatorname{M\ddot{o}b} (and we recall α(γ)(z)=1/γ(1z)\alpha(\gamma)(z)=1/\gamma(\tfrac{1}{z}))444It may be surprising that the condition gLn=L~nggL_{n}=\tilde{L}_{n}g for vertex algebras corresponds to gU(γ)=U~(α(γ))ggU(\gamma)=\tilde{U}(\alpha(\gamma))g for Wightman CFTs. Note that due to the antilinearity of gg, the relation gLn=L~nggL_{n}=\tilde{L}_{n}g does not imply that gg intertwines the representations of Lie(Mo¨b)\operatorname{Lie}(\operatorname{M\ddot{o}b}). In fact we have gπ(X)=π~(dα(X))gg\pi(X)=\tilde{\pi}(d\alpha(X))g for XLie(Mo¨b)X\in\operatorname{Lie}(\operatorname{M\ddot{o}b}).. We have

U~(γ)(gφ)(f)=gU(α(γ))φ(f¯1z),\tilde{U}(\gamma)(g_{*}\varphi)(f)=gU(\alpha(\gamma))\varphi(\overline{f}\circ\tfrac{1}{z}),

where f¯\overline{f} denotes the pointwise complex conjugate. Just as we demonstrated in Lemmas 3.15 and 3.16, one can show that antilinear homomorphisms of Möbius vertex algebras extend uniquely to antilinear homomorphisms of Wightman CFTs, and conversely antilinear homomorphisms of Wightman CFTs restrict to antilinear homomorphisms of Möbius vertex algebras.

Lemma 4.6.

Let (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) and (~,𝒟~,U~,Ω~)(\tilde{\mathcal{F}},\tilde{\mathcal{D}},\tilde{U},\tilde{\Omega}) be two Möbius-covariant Wightman CFTs and let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} and 𝒱~𝒟~\tilde{\mathcal{V}}\subset\tilde{\mathcal{D}} be the corresponding Möbius vertex algebras with respective generating sets SS and S~\tilde{S}.

  1. i)

    If (g,g):~(g,g_{*}):{\mathcal{F}}\to\tilde{\mathcal{F}} is an antilinear homomorphism then g(𝒱)𝒱~g({\mathcal{V}})\subset\tilde{\mathcal{V}} and g|𝒱g|_{{\mathcal{V}}} is an antilinear homomorphism of Möbius vertex algebras satisfying g(S)S~g(S)\subset\tilde{S}.

  2. ii)

    If g:𝒱𝒱~g:{\mathcal{V}}\to\tilde{\mathcal{V}} is an antilinear homomorphism of Möbius vertex algebras such that g(S)S~g(S)\subset\tilde{S}, then there is a unique antilinear homomorphism (h,h):~(h,h_{*}):{\mathcal{F}}\to\tilde{\mathcal{F}} such that h|𝒱=gh|_{\mathcal{V}}=g.

We omit the proof of Lemma 4.6 which is essentially identical to Lemma 3.15 and 3.16 (once we observe as above that an antilinear vertex algebra homomorphism satisfies gπ(X)=π~(dα(X))gg\pi(X)=\tilde{\pi}(d\alpha(X))g for XLie(Mo¨b)X\in\operatorname{Lie}(\operatorname{M\ddot{o}b})).

Recall that an antilinear map gg is said to preserve a sesquilinear form if gΦ,gΨ=Φ,Ψ¯\langle g\Phi,g\Psi\rangle=\overline{\langle\Phi,\Psi\rangle} for all vectors Φ,Ψ\Phi,\Psi, and that a sesquilinear form is said to be (Hermitian) symmetric if Φ,Ψ=Ψ,Φ¯\langle\Phi,\Psi\rangle=\overline{\langle\Psi,\Phi\rangle}.

Definition 4.7.

An involutive Möbius vertex algebra is a Möbius vertex algebra 𝒱{\mathcal{V}} equipped with a sesquilinear form ,\langle\,\cdot\,,\,\cdot\,\rangle and an antilinear automorphism θ:𝒱𝒱\theta:{\mathcal{V}}\to{\mathcal{V}} which is involutive (θ2=id𝒱)(\theta^{2}=\mathrm{id}_{\mathcal{V}}) and preserves the sesquilinear form, and such that ,θ\langle\,\cdot\,,\theta\,\cdot\,\rangle is an invariant bilinear form. An involutive Möbius vertex algebra is called unitary if the sesquilinear form is an inner product that is normalized so that Ω,Ω=1\langle\Omega,\Omega\rangle=1.

We use the convention that sesquilinear forms are linear in the first variable, and require that homomorphisms of Möbius vertex algebras commute with the operators LnL_{n}. The condition that ,θ\langle\,\cdot\,,\theta\,\cdot\,\rangle is an invariant bilinear form is equivalent to having

Y(v,z)u1,u2=u1,Y𝗈(θv,z¯)u2 and Lnu1,u2=u1,Lnu2\langle Y(v,z)u_{1},u_{2}\rangle=\langle u_{1},Y^{\mathsf{o}}(\theta v,\bar{z})u_{2}\rangle\quad\mbox{ and }\quad\langle L_{n}u_{1},u_{2}\rangle=\langle u_{1},L_{-n}u_{2}\rangle (4.9)

for all u1,u2,v𝒱u_{1},u_{2},v\in{\mathcal{V}} and n=1,0,1n=-1,0,1, where zz is a formal complex variable, i.e.​ ,z¯=z,\langle\,\cdot\,,\bar{z}\,\cdot\,\rangle=z\langle\,\cdot\,,\cdot\,\rangle.

We sometimes refer to the sesquilinear form from Definition 4.7 as an invariant sesquilinear form, omitting reference to the involution θ\theta.

Remark 4.8.

The sesquilinear forms from Definition 4.7 are automatically Hermitian symmetric as a consequence of the fact that invariant bilinear forms are symmetric. If the sesquilinear form is nondegenerate then the requirement that θ\theta be involutive is redundant and the automorphism θ\theta is uniquely determined by the sesquilinear form (the proof is exactly as in [CKLW18, Prop. 5.1] for inner products).

We now turn our attention to invariant sesquilinear forms on Wightman CFTs.

Definition 4.9.

An involutive Möbius-covariant Wightman CFT on S1S^{1} is a Wightman CFT (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) along with a jointly {\mathcal{F}}-strong continuous sesquilinear form ,\langle\,\cdot\,,\,\cdot\,\rangle on 𝒟{\mathcal{D}} and an involutive automorphism (θ,θ)(\theta,\theta_{*}) of {\mathcal{F}} such that θ\theta preserves the sesquilinear form and such that ,θ\langle\,\cdot\,,\,\theta\cdot\,\rangle is an invariant bilinear form. An involutive Möbius-covariant Wightman CFT is called unitary if the sesquilinear form is an inner product which is normalized so that Ω,Ω=1\langle\Omega,\Omega\rangle=1.

As with vertex algebras, we sometimes refer to the sesquilinear form of Definition 4.9 as an invariant sesquilinear form, omitting reference to the involution.

If we write φ=(1)dφθφ\varphi^{\dagger}=(-1)^{d_{\varphi}}\theta_{*}\varphi, then the condition that ,θ\langle\,\cdot\,,\,\theta\cdot\,\rangle is an invariant bilinear form is equivalent to

φ(f)Φ,Ψ=Φ,φ(f¯)Ψ and U(γ)Φ,U(γ)Ψ=Φ,Ψ\langle\varphi(f)\Phi,\Psi\rangle=\langle\Phi,\varphi^{\dagger}(\overline{f})\Psi\rangle\quad\mbox{ and }\quad\langle U(\gamma)\Phi,U(\gamma)\Psi\rangle=\langle\Phi,\Psi\rangle (4.10)

for all Φ,Ψ𝒟\Phi,\Psi\in{\mathcal{D}}, φ\varphi\in{\mathcal{F}}, and γMo¨b\gamma\in\operatorname{M\ddot{o}b}555Note that this is a slight departure from [RTT22], where we required that {\mathcal{F}} be invariant under the involution \dagger rather than θ\theta_{*}. In the present setting we find the updated definition to be more natural, as the involution \dagger of fields typically does not correspond to an antilinear automorphism of the Wightman CFT.. Here, as before, f¯\overline{f} denotes the pointwise complex conjugate of the function ff.

As with involutive vertex algebras (Remark 4.8), the sesquilinear form of an involutive Wightman CFT is automatically Hermitian symmetric.

Theorem 4.10 (Equivalence of involutive and unitary structures).

Let (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) be a Möbius-covariant Wightman CFT, and let 𝒱𝒟{\mathcal{V}}\subset{\mathcal{D}} be the corresponding Möbius vertex algebra equipped with a set SS of quasiprimary generators. Then we have the following.

  1. i)

    If 𝒟{\mathcal{D}} is equipped with a sesquilinear form and involution (θ,θ)(\theta,\theta_{*}) making it into an involutive Wightman CFT, then the sesquilinear form and involution θ\theta restrict to an involutive structure on the vertex algebra 𝒱{\mathcal{V}}. The set S𝒱S\subset{\mathcal{V}} of quasiprimary generators is invariant under θ\theta.

  2. ii)

    If 𝒱{\mathcal{V}} is equipped with a sesquilinear form and involution θ\theta making it into an involutive vertex algebra and SS is invariant under θ\theta, then there is a unique involution θ\theta_{*} of {\mathcal{F}} and unique extensions of the sesquilinear form and θ\theta to 𝒟{\mathcal{D}} making {\mathcal{F}} into an involutive Wightman CFT.

If the sesquilinear form is nondegenerate on 𝒟{\mathcal{D}} then it remains nondegenerate on 𝒱{\mathcal{V}}, and similarly if the form is nondegenerate on 𝒱{\mathcal{V}} so is the extension to 𝒟{\mathcal{D}}. Moreover unitary structures on 𝒟{\mathcal{D}} correspond to unitary structures on 𝒱{\mathcal{V}}, and vice versa.

Proof.

First consider an involutive structure (θ,θ)(\theta,\theta_{*}) on {\mathcal{F}}. Then θ\theta restricts to an antilinear involutive Möbius vertex algebra automorphism θ|𝒱:𝒱𝒱\theta|_{\mathcal{V}}:{\mathcal{V}}\to{\mathcal{V}} by Lemma 4.6. By Theorem 4.3, the invariant bilinear form ,θ\langle\,\cdot\,,\,\theta\cdot\,\rangle restricts to an invariant bilinear form on 𝒱{\mathcal{V}}, and it follows that ,\langle\,\cdot\,,\,\cdot\,\rangle and θ|𝒱\theta|_{{\mathcal{V}}} yield an involutive structure on 𝒱{\mathcal{V}}. If φ\varphi\in{\mathcal{F}} corresponds to the state v𝒱(d)v\in{\mathcal{V}}(d), then

θv=θφ(ed)Ω=(θφ)(ed)Ω.\theta v=\theta\varphi(e_{-d})\Omega=(\theta_{*}\varphi)(e_{-d})\Omega.

Hence the Wightman field θφ\theta_{*}\varphi\in{\mathcal{F}} corresponds to θv\theta v and SS is θ\theta-invariant (we have used here the observation that θ\theta_{*} preserves the conformal dimension of fields).

For the other direction, suppose that we have an involutive structure on 𝒱{\mathcal{V}} corresponding to an involution θ\theta and sesquilinear form ,\langle\,\cdot\,,\,\cdot\,\rangle. Then the invariant bilinear form ,θ\langle\,\cdot\,,\theta\,\cdot\,\rangle extends uniquely to an invariant bilinear form (,)(\,\cdot\,,\,\cdot\,) on 𝒟{\mathcal{D}}. By Lemma 4.6 we may uniquely extend θ\theta to an antilinear automorphism (θ,θ)(\theta,\theta_{*}) of {\mathcal{F}}. This extension is {\mathcal{F}}-strong continuous, and thus the sesquilinear form (,θ)(\,\cdot\,,\,\theta\cdot\,) is {\mathcal{F}}-strong continuous as well. This sesquilinear form, along with (θ,θ)(\theta,\theta_{*}), yield an involutive structure on {\mathcal{F}} as required.

The proof of equivalence of nondegeneracy is straightforward (as in the proof of Theorem 4.3), and the equivalence of unitarity is immediate. ∎

For unitary Wightman CFTs domain 𝒟{\mathcal{D}} can be equipped with the norm topology coming from the inner product. This leads to a number of analytic questions, which are discussed further in [RTT22].

Remark 4.11.

Suppose that (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) is a Wightman CFT such that 𝒟{\mathcal{D}} is equipped with an inner product and {\mathcal{F}} is equipped with an involution \dagger. The definition of a (unitary) Wightman CFT given in [RTT22] required only that the compatibility conditions (4.10) hold, with no mention of the PCT operator θ\theta. However, under these assumptions one may show that there exists a unique θ\theta making the associated vertex algebra 𝒱{\mathcal{V}} into a unitary vertex algebra, arguing as in [RTT22, Thm. 3.11] based on [CKLW18, Thm. 5.16]666The hypothesis that dim𝒱(0)=1\dim{\mathcal{V}}(0)=1 required in [CKLW18, Thm. 5.16] is not needed, as shown in [CGH23, §3.4], and the condition dim𝒱(n)<\dim{\mathcal{V}}(n)<\infty is also not needed.. One may then extend θ\theta to an antilinear involution (θ,θ)(\theta,\theta_{*}) of {\mathcal{F}} by Lemma 4.6, making (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) into a unitary Wightman CFT as defined in this article. We note that for a general sesquilinear form on 𝒟{\mathcal{D}}, an involution \dagger satisfying (4.10) does not necessarily correspond to an involutive structure. Indeed, in the extreme example where the sesquilinear form is identically zero, the compatibility conditions (4.10) impose no constraint on the involution \dagger, but not every set-theoretic involution of {\mathcal{F}} corresponds to an involutive structure. It is possible that the conditions (4.10) are sufficient to reconstruct the PCT operator θ\theta when the sesquilinear form is nondegenerate, but we do not address that question here.

Appendix A The Reeh-Schlieder theorem for non-unitary
Wightman conformal field theories

In this section we work with rotation-covariant Wightman CFTs (,𝒟,U,Ω)({\mathcal{F}},{\mathcal{D}},U,\Omega) on S1S^{1}, which differ from Möbius-covariant Wightman CFTs (Definition 2.9) only in that the symmetry UU is only a representation of of the rotation subgroup Rot(S1)Mo¨b\operatorname{Rot}(S^{1})\subset\operatorname{M\ddot{o}b}, and accordingly the covariance condition (W1) is weakened to only require covariance for rotations. We write either RzR_{z} or RϑR_{\vartheta} for rotation by z=eiϑz={\mathrm{e}}^{{\mathrm{i}}\vartheta}.

For IS1I\subset S^{1} an interval (i.e.​ II is a connected open non-empty proper subset), we let 𝒫(I)(𝒟){\mathcal{P}}(I)\subset{\mathcal{L}}({\mathcal{D}}) be the algebra generated by smeared fields φ(f)\varphi(f) with suppfI\operatorname{supp}f\subset I. The goal of this section is to establish the Reeh-Schlieder theorem for the theory {\mathcal{F}}, which says that the vacuum vector Ω\Omega is cyclic and separating for the algebras 𝒫(I){\mathcal{P}}(I). Recall that Ω\Omega is cyclic for an algebra 𝒫(𝒟){\mathcal{P}}\subset{\mathcal{L}}({\mathcal{D}}) (with respect to a certain topology on 𝒟{\mathcal{D}}) if 𝒫Ω{\mathcal{P}}\Omega is dense in 𝒟{\mathcal{D}}, and separating for 𝒫{\mathcal{P}} if the only X𝒫X\in{\mathcal{P}} such that XΩ=0X\Omega=0 is X=0X=0. The analogous statement for unitary Wightman quantum field theories on higher-dimensional spacetimes is well-known (see [SW64, §4.2] and [RS61]). We give here a proof of the Reeh-Schlieder theorem in our current (not necessarily unitary) context.

Let DD be the open unit disk in {\mathbb{C}}, and D¯\overline{D} its closure. We denote by A(D¯)A(\overline{D}) the space of continuous {\mathbb{C}}-valued functions on D¯\overline{D} that are holomorphic on the interior DD. By the maximum principle A(D¯)A(\overline{D}) embeds as a closed subspace of C(S1)C(S^{1}), and we give A(D¯)A(\overline{D}) the norm inherited from C(S1)C(S^{1}).

Lemma A.1.

Let {\mathcal{F}} be a rotation-covariant Wightman CFT. Fix φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}}, and let λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}. Let f1,,fkC(S1)f_{1},\ldots,f_{k}\in C^{\infty}(S^{1}) and let z1,,zkS1z_{1},\ldots,z_{k}\in S^{1}. Then for each j=1,,kj=1,\cdots,k, the map

zjλ(U(Rz1)φ1(f1)U(Rz2)φ2(f2)U(Rzk)φk(fk)Ω)z_{j}\mapsto\lambda\big{(}U(R_{z_{1}})\varphi_{1}(f_{1})U(R_{z_{2}})\varphi_{2}(f_{2})\cdots U(R_{z_{k}})\varphi_{k}(f_{k})\Omega\big{)} (A.1)

lies in A(D¯)A(\overline{D}).

Proof.

When the functions fjf_{j} are all Laurent polynomials, the expression (A.1) is a polynomial in the ziz_{i} and the conclusion follows. We now consider the general case.

By rotation covariance we have

λ(U(Rz1)φ1(f1)U(Rz2)U(Rzk)φk(fk)Ω)=λ(φ1(βd1(Rw1)f1)φk(βdk(Rwk)fk)Ω)\lambda\big{(}U(R_{z_{1}})\varphi_{1}(f_{1})U(R_{z_{2}})\cdots U(R_{z_{k}})\varphi_{k}(f_{k})\Omega\big{)}=\lambda\big{(}\varphi_{1}(\beta_{d_{1}}(R_{w_{1}})f_{1})\cdots\varphi_{k}(\beta_{d_{k}}(R_{w_{k}})f_{k})\Omega\big{)}

where wj=z1z2zjw_{j}=z_{1}z_{2}\cdots z_{j}, and djd_{j} is the conformal dimension of φj\varphi_{j}. Given arbitrary smooth fjf_{j}, choose sequences of Laurent polynomials fj,nf_{j,n} such that fj,nfjf_{j,n}\to f_{j} in C(S1)C^{\infty}(S^{1}). As in Section 3.1, let HN(S1)H^{N}(S^{1}) be the Sobolev space corresponding to a number N>0N>0, and recall that the topology on C(S1)C^{\infty}(S^{1}) is generated by the Sobolev norms N\left\|\,\cdot\,\right\|_{N}. Since βd(Rw)\beta_{d}(R_{w}) acts as a unitary on HN(S1)H^{N}(S^{1}), we have convergence in each HN(S1)H^{N}(S^{1})

limnβdj(Rw)fj,n=βdj(Rw)fj\lim_{n\to\infty}\beta_{d_{j}}(R_{w})f_{j,n}=\beta_{d_{j}}(R_{w})f_{j}

that is uniform in ww.

Since λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}, expressions

λ(φ1(f1)φk(fk)Ω)\lambda\big{(}\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Omega\big{)} (A.2)

are jointly continuous as maps C(S1)kC^{\infty}(S^{1})^{k}\to{\mathbb{C}}. Hence we may choose a positive number NN such that (A.2) is jointly continuous HN(S1)kH^{N}(S^{1})^{k}\to{\mathbb{C}} (i.e.​ it is a bounded multilinear map). It follows that

limnλ(φ1(βd1(Rw1)f1,n)φk(βdk(Rwk)fk,n)Ω)=λ(φ1(βd1(Rw1)f1)φk(βdk(Rwk)fk))\lim_{n\to\infty}\lambda\big{(}\varphi_{1}(\beta_{d_{1}}(R_{w_{1}})f_{1,n})\cdots\varphi_{k}(\beta_{d_{k}}(R_{w_{k}})f_{k,n})\Omega\big{)}=\lambda\big{(}\varphi_{1}(\beta_{d_{1}}(R_{w_{1}})f_{1})\cdots\varphi_{k}(\beta_{d_{k}}(R_{w_{k}})f_{k})\big{)}

uniformly in z1,,zkz_{1},\ldots,z_{k}. As each map

zj\displaystyle z_{j} λ(U(Rz1)φ1(f1,n)U(Rz2)φ2(f2)U(Rzk)φk(fk,n)Ω)\displaystyle\mapsto\lambda\big{(}U(R_{z_{1}})\varphi_{1}(f_{1,n})U(R_{z_{2}})\varphi_{2}(f_{2})\cdots U(R_{z_{k}})\varphi_{k}(f_{k,n})\Omega\big{)}
=λ(φ1(βd1(Rw1)f1,n)φk(βdk(Rwk)fk,n)Ω)\displaystyle=\lambda\big{(}\varphi_{1}(\beta_{d_{1}}(R_{w_{1}})f_{1,n})\cdots\varphi_{k}(\beta_{d_{k}}(R_{w_{k}})f_{k,n})\Omega\big{)}

lies in A(D¯)A(\overline{D}) and A(D¯)A(\overline{D}) is a closed subspace of C(S1)C(S^{1}), the map (A.1) lies in A(D¯)A(\overline{D}) as claimed. ∎

Lemma A.2.

Let {\mathcal{F}} be a rotation-covariant Wightman CFT on S1S^{1} with domain 𝒟{\mathcal{D}}, and let IS1I\subset S^{1} be an interval. Let λ𝒟\lambda\in{\mathcal{D}}_{\mathcal{F}}^{*}, and suppose λ(XΩ)=0\lambda(X\Omega)=0 for all X𝒫(I)X\in{\mathcal{P}}(I). Then λ=0\lambda=0.

Proof.

Fix φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}}, so that

λ(φ1(f1)φk(fn)Ω)=0\lambda\big{(}\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{n})\Omega\big{)}=0 (A.3)

whenever supp(fj)I\operatorname{supp}(f_{j})\subset I for j=1,,kj=1,\ldots,k. Fix f1,,fkf_{1},\ldots,f_{k} supported in II, and consider the function Fk:S1F_{k}:S^{1}\to{\mathbb{C}} given by

Fk(z)=λ(φ1(f1)φk1(fk1)U(Rz)φk(fk)Ω).F_{k}(z)=\lambda\big{(}\varphi_{1}(f_{1})\cdots\varphi_{k-1}(f_{k-1})U(R_{z})\varphi_{k}(f_{k})\Omega\big{)}.

We have FkA(D¯)F_{k}\in A(\overline{D}) by Lemma A.1. Moreover, by rotation covariance FkF_{k} vanishes on a small interval of S1S^{1} about 11 (note that supp(f)\operatorname{supp}(f) is closed and the interval II is open, so that II contains a neighborhood of supp(fk)\operatorname{supp}(f_{k})). Thus by the Schwarz reflection principle we have Fk=0F_{k}=0 identically, and restricting to zS1z\in S^{1} we have

0=Fk(z)=λ(φ1(f1)φk1(fk1)φk(βd(Rz)fk)Ω)0=F_{k}(z)=\lambda\big{(}\varphi_{1}(f_{1})\cdots\varphi_{k-1}(f_{k-1})\varphi_{k}(\beta_{d}(R_{z})f_{k})\Omega\big{)}

for all zS1z\in S^{1}. Hence (A.3) holds whenever f1,,fk1f_{1},\ldots,f_{k-1} are supported in II, and fkf_{k} is supported in any interval of length |I|\left|I\right|. Using a partition of unity, it follows that (A.3) holds for arbitrary fkf_{k}.

We now repeat the above argument. As before, we may show that the function

zλ(φ1(f1)U(Rz)φk1(fk1)φk(fk)Ω)z\mapsto\lambda\big{(}\varphi_{1}(f_{1})\cdots U(R_{z})\varphi_{k-1}(f_{k-1})\varphi_{k}(f_{k})\Omega\big{)}

vanishes identically on S1S^{1}, and from there deduce that (A.3) holds whenever f1,,fk2f_{1},\ldots,f_{k-2} are supported in II, and fk1,fkf_{k-1},f_{k} are arbitrary. Repeatedly applying this argument, we see that (A.3) holds for all f1,,fkC(S1)f_{1},\ldots,f_{k}\in C^{\infty}(S^{1}), which means λ=0\lambda=0 by the vacuum axiom of a Wightman CFT. ∎

Corollary A.3 (Reeh-Schlieder theorem).

Let {\mathcal{F}} be a rotation-covariant Wightman CFT on S1S^{1} with domain 𝒟{\mathcal{D}}. For IS1I\subset S^{1} an interval we let 𝒫(I)(𝒟){\mathcal{P}}(I)\subset{\mathcal{L}}({\mathcal{D}}) be the subalgebra generated by φ(f)\varphi(f) with φ\varphi\in{\mathcal{F}} and supp(f)I\operatorname{supp}(f)\subset I. Then

  1. i)

    Ω\Omega is cyclic for 𝒫(I){\mathcal{P}}(I) with respect to the {\mathcal{F}}-strong topology on 𝒟{\mathcal{D}}, i.e.​ 𝒫(I)Ω{\mathcal{P}}(I)\Omega is {\mathcal{F}}-strongly dense in 𝒟{\mathcal{D}}.

  2. ii)

    Ω\Omega is separating for 𝒫(I){\mathcal{P}}(I), i.e.​ if X𝒫(I)X\in{\mathcal{P}}(I) and XΩ=0X\Omega=0 then X=0X=0.

Proof.

For part (i), recall from Remark 2.6 that 𝒟{\mathcal{D}}_{\mathcal{F}}^{*} is precisely the dual space of 𝒟{\mathcal{D}} equipped with the {\mathcal{F}}-strong topology. By Lemma A.2 the closed subspace 𝒫(I)Ω¯\overline{{\mathcal{P}}(I)\Omega} is annihilated only by the zero functional, and so by the Hahn-Banach theorem (for locally convex topological vector spaces) we must have 𝒫(I)Ω¯=𝒟\overline{{\mathcal{P}}(I)\Omega}={\mathcal{D}}.

For part (ii), observe that by the locality axiom of a Wightman theory the operator XX vanishes on 𝒫(I)Ω{\mathcal{P}}(I^{\prime})\Omega, where II^{\prime} is the interval complementary to II. By Lemma 2.8 the operator X:𝒟𝒟X:{\mathcal{D}}\to{\mathcal{D}} is {\mathcal{F}}-strongly continuous, and hence by part (i) we have X=0X=0. ∎

Appendix B Topological vector spaces

In this section we supplement the discussion of the topology on the domain 𝒟{\mathcal{D}} of a Wightman field theory by giving additional definitions, details, and references regarding topological vector spaces and locally convex spaces. We refer readers to the textbooks [NB11, Trè67] for further reading. All vector spaces in this section are assumed to be over the field of complex numbers.

A topological vector space is a vector space VV equipped with a vector topology, which is a topology such that the addition map V×VVV\times V\to V and the scalar multiplication map ×VV{\mathbb{C}}\times V\to V are continuous. Vector topologies are not necessarily Hausdorff by definition, although we will primarily be interested in Hausdorff topological vector spaces.

A seminorm on a vector space VV is a map p:V0p:V\to{\mathbb{R}}_{\geq 0} such that p(u+v)p(u)+p(v)p(u+v)\leq p(u)+p(v) and p(αu)=|α|p(u)p(\alpha u)=\left|\alpha\right|p(u) for all u,vVu,v\in V and α\alpha\in{\mathbb{C}}. Given a set of seminorms on VV, the corresponding seminorm topology is the coarsest topology on VV making all of the seminorms continuous. Seminorm topologies are always vector topologies, but not every vector topology is a seminorm topology. A locally convex space is a topological vector space whose topology is a seminorm topology corresponding to some set of seminorms. Equivalently, a locally convex space is a topological vector space such that there exists a neighborhood basis of the origin consisting of convex sets [NB11, Thm. 5.5.2]. Every Hausdorff topological vector space VV has a unique completion V^\widehat{V} [Trè67, §5], and the completion of a locally convex space is locally convex [NB11, Thm. 5.11.5]. We note that finite-dimensional Hausdorff topological vector spaces are complete [NB11, Thm. 4.10.3], as are products of complete topological vector spaces. Every continuous linear map T:UVT:U\to V of Hausdorff topological vector spaces extends continuously to a map T^:U^V^\widehat{T}:\widehat{U}\to\widehat{V} [Trè67, Thm. 5.2].

Locally convex spaces play an important role in functional analysis because the Hahn-Banach theorem holds for them. In particular, the continuous linear functionals on a locally convex Hausdorff space separate points. Moreover, if XX is a closed subspace of a locally convex Hausdorff space VV and vXv\not\in X, then there exists a continuous linear functional λ:V\lambda:V\to{\mathbb{C}} such that λ|X0\lambda|_{X}\equiv 0 and λ(v)=1\lambda(v)=1 [NB11, Thm. 7.7.7]. In contrast, there exist topological vector spaces which do not admit nonzero continuous linear functionals, such as LpL^{p} spaces with 0<p<10<p<1.

Most familiar examples of topological vector spaces, such as normed vector spaces, are locally convex. Another source of locally convex spaces is via weak topologies [NB11, §8.2]. Given a vector space VV and a set of linear functionals 𝒳{\mathcal{X}} on VV, the weak topology (or initial topology) on VV corresponding to 𝒳{\mathcal{X}} is the coarsest topology making all of the functionals continuous. This is a locally convex vector topology, being the seminorm topology corresponding to the seminorms |λ|\left|\lambda\right| for λ𝒳\lambda\in{\mathcal{X}}. A sequence (or net) vnVv_{n}\in V converges to vv if and only if λ(vn)λ(v)\lambda(v_{n})\to\lambda(v) for every λ𝒳\lambda\in{\mathcal{X}}. A map T:XVT:X\to V is continuous with respect to the weak topology if and only if λT\lambda\circ T is continuous for every λ𝒳\lambda\in{\mathcal{X}}.

Dually, we have the notion of the colimit (or final or strong) topology. Consider a vector space VV, and a family of linear maps Ts:XsVT_{s}:X_{s}\to V from topological vector spaces XsX_{s} such that the images Ts(Xs)T_{s}(X_{s}) span VV. The colimit topology on VV corresponding to the maps TsT_{s} is the finest topology on VV such that every TsT_{s} is continuous, and it is a vector topology [NB11, §4.11]. If UU is a topological vector space, then a linear map T:VUT:V\to U is continuous if and only if TTsT\circ T_{s} is continuous for all ss.

If each space XsX_{s} is locally convex then we may define a subtly different locally convex colimit topology on VV, which is the finest locally convex topology such that each XsX_{s} is continuous [NB11, §12.2]. If UU is a locally convex space then a linear map T:VUT:V\to U is continuous for the locally convex colimit topology if and only if TTsT\circ T_{s} is continuous for all ss [NB11, Thm. 12.2.2].

We now discuss tensor products of locally convex spaces. If UU,VV, and XX are vector spaces then bilinear maps U×VXU\times V\to X correspond to linear maps UVXU\otimes V\to X, where \otimes is the algebraic tensor product. If UU,VV, and XX are locally convex spaces, then there is a unique locally convex topology on UVU\otimes V, called the π\pi-topology (or projective topology), such that jointly continuous bilinear maps U×VXU\times V\to X correspond to continuous linear maps UVXU\otimes V\to X [Trè67, Prop. 43.4]. We write UπVU\otimes_{\pi}V for the algebraic tensor product equipped with the π\pi topology.

We now conclude by revisiting the {\mathcal{F}}-strong topology. Suppose that {\mathcal{F}} is a set of operator-valued distributions on S1S^{1} with domain a vector space 𝒟{\mathcal{D}}. For every φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}} we have a multilinear map C(S1)k𝒟C^{\infty}(S^{1})^{k}\to{\mathcal{D}} given by (f1,,fk)φ1(f1)φk(fk)Φ(f_{1},\ldots,f_{k})\mapsto\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi. These correspond to linear maps

Sφ1,,φk,Φ:C(S1)ππC(S1)𝒟.S_{\varphi_{1},\ldots,\varphi_{k},\Phi}:C^{\infty}(S^{1})\otimes_{\pi}\cdots\otimes_{\pi}C^{\infty}(S^{1})\to{\mathcal{D}}.

We include the case k=0k=0, in which case SΦ:𝒟S_{\Phi}:{\mathbb{C}}\to{\mathcal{D}} assigns 1Φ1\mapsto\Phi. The {\mathcal{F}}-strong topology on 𝒟{\mathcal{D}} is then defined to be the locally convex colimit of the maps Sφ1,,φk,ΦS_{\varphi_{1},\ldots,\varphi_{k},\Phi}. Unpacking the definitions, if XX is a locally convex space then a map T:𝒟XT:{\mathcal{D}}\to X is {\mathcal{F}}-strong continuous if and only if T(φ1(f1)φk(fk)Φ)T(\varphi_{1}(f_{1})\cdots\varphi_{k}(f_{k})\Phi) is jointly continuous in the fjf_{j} for all φ1,,φk\varphi_{1},\ldots,\varphi_{k}\in{\mathcal{F}} and Φ𝒟\Phi\in{\mathcal{D}}.

References

  • [AGT23] M. S. Adamo, L. Giorgetti, and Y. Tanimoto. Wightman fields for two-dimensional conformal field theories with pointed representation category. Comm. Math. Phys., 404(3):1231–1273, 2023.
  • [AMT24] M. S. Adamo, Y. Moriwaki, and Y. Tanimoto. Osterwalder–Schrader axioms for unitary full vertex operator algebras. arXiv:2407.18222 [math-ph], 2024.
  • [BK08] G. Buhl and G. Karaali. Spanning sets for Möbius vertex algebras satisfying arbitrary difference conditions. J. Algebra, 320(8):3345–3364, 2008.
  • [CGH23] S. Carpi, T. Gaudio, and R. Hillier. From vertex operator superalgebras to graded-local conformal nets and back. arXiv:2304.14263 [math.OA], 2023.
  • [CKLW18] S. Carpi, Y. Kawahigashi, R. Longo, and M. Weiner. From vertex operator algebras to conformal nets and back. Mem. Amer. Math. Soc., 254(1213), 2018.
  • [CWX] S. Carpi, M. Weiner, and F. Xu. From vertex operator algebra modules to representations of conformal nets. In preparation.
  • [EG17] D. E. Evans and T. Gannon. Non-unitary fusion categories and their doubles via endomorphisms. Adv. Math., 310:1–43, 2017.
  • [FHL93] I. B. Frenkel, Y.-Z. Huang, and J. Lepowsky. On axiomatic approaches to vertex operator algebras and modules. Mem. Amer. Math. Soc., 104(494):viii+64, 1993.
  • [Gui20] B. Gui. Unbounded field operators in categorical extensions of conformal nets. arXiv:2001.03095 [math.QA], 2020.
  • [Gui21] B. Gui. Categorical extensions of conformal nets. Comm. Math. Phys., 383(2):763–839, 2021.
  • [Gui24a] B. Gui. Convergence of sewing conformal blocks. Commun. Contemp. Math., 26(3):Paper No. 2350007, 65, 2024.
  • [Gui24b] B. Gui. Sewing and propagation of conformal blocks. New York J. Math., 30:187–230, 2024.
  • [GZ23] B. Gui and H. Zhang. Analytic Conformal Blocks of C2C_{2}-cofinite Vertex Operator Algebras I: Propagation and Dual Fusion Products. arXiv:2305.10180 [math.QA], 2023.
  • [HLZ14] Y.-Z. Huang, J. Lepowsky, and L. Zhang. Logarithmic tensor category theory for generalized modules for a conformal vertex algebra, I: introduction and strongly graded algebras and their generalized modules. In Conformal field theories and tensor categories, Math. Lect. Peking Univ., pages 169–248. Springer, Heidelberg, 2014.
  • [Hua99] Y.-Z. Huang. A functional-analytic theory of vertex (operator) algebras. I. Comm. Math. Phys., 204(1):61–84, 1999.
  • [Hua03] Y.-Z. Huang. A functional-analytic theory of vertex (operator) algebras. II. Comm. Math. Phys., 242(3):425–444, 2003.
  • [Hua20] Y.-Z. Huang. Generators, spanning sets and existence of twisted modules for a grading-restricted vertex (super)algebra. Selecta Math. (N.S.), 26(4):Paper No. 62, 42, 2020.
  • [Kac98] V. Kac. Vertex algebras for beginners, volume 10 of University Lecture Series. American Mathematical Society, Providence, RI, second edition, 1998.
  • [Li94] H. S. Li. Symmetric invariant bilinear forms on vertex operator algebras. J. Pure Appl. Algebra, 96(3):279–297, 1994.
  • [Mor23] Y. Moriwaki. Two-dimensional conformal field theory, full vertex algebra and current-current deformation. Adv. Math., 427:Paper No. 109125, 74, 2023.
  • [NB11] L. Narici and E. Beckenstein. Topological vector spaces, volume 296 of Pure and Applied Mathematics (Boca Raton). CRC Press, Boca Raton, FL, second edition, 2011.
  • [RS61] H. Reeh and S. Schlieder. Bemerkungen zur Unitäräquivalenz von Lorentzinvarianten Felden. Nuovo Cimento (10), 22:1051–1068, 1961.
  • [RTT22] C. Raymond, Y. Tanimoto, and J. E. Tener. Unitary vertex algebras and Wightman conformal field theories. Comm. Math. Phys., 395(1):299–330, 2022.
  • [Seg04] G. Segal. The definition of conformal field theory. In Topology, geometry and quantum field theory, volume 308 of London Math. Soc. Lecture Note Ser., pages 421–577. Cambridge Univ. Press, Cambridge, 2004.
  • [Str93] F. Strocchi. Selected topics on the general properties of quantum field theory, volume 51 of World Scientific Lecture Notes in Physics. World Scientific Publishing Co., Inc., River Edge, NJ, 1993.
  • [SW64] R. F. Streater and A. S. Wightman. PCT, spin and statistics, and all that. W. A. Benjamin, Inc., New York-Amsterdam, 1964.
  • [Ten19a] J. E. Tener. Geometric realization of algebraic conformal field theories. Adv. Math., 349:488–563, 2019.
  • [Ten19b] J. E. Tener. Representation theory in chiral conformal field theory: from fields to observables. Selecta Math. (N.S.), 25(5):Paper No. 76, 82, 2019.
  • [Ten24] J. E. Tener. Fusion and positivity in chiral conformal field theory. Geom. Funct. Anal., 34(4):1226–1296, 2024.
  • [Trè67] F. Trèves. Topological vector spaces, distributions and kernels. Academic Press, New York-London, 1967.